Accepted Manuscript Parametric analysis and optimization of an Organic Rankine Cycle with nanofluid based solar parabolic trough collectors Evangelos Bellos, Christos Tzivanidis PII:
S0960-1481(17)30557-8
DOI:
10.1016/j.renene.2017.06.055
Reference:
RENE 8922
To appear in:
Renewable Energy
Received Date: 10 May 2017 Revised Date:
31 May 2017
Accepted Date: 14 June 2017
Please cite this article as: Bellos E, Tzivanidis C, Parametric analysis and optimization of an Organic Rankine Cycle with nanofluid based solar parabolic trough collectors, Renewable Energy (2017), doi: 10.1016/j.renene.2017.06.055. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. 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.
ACCEPTED MANUSCRIPT
Parametric analysis and optimization of an Organic Rankine Cycle
2
with nanofluid based solar parabolic trough collectors
3
Evangelos Bellos, Christos Tzivanidis
4
Thermal Department, School of Mechanical Engineering, National Technical
5
University of Athens, Zografou, Heroon Polytechniou 9, 15780 Athens, Greece.
RI PT
1
6
Corresponding author: Evangelos Bellos (
[email protected], tel:+302107722340)
SC
7
Abstract
9
The main objective of this work is to investigate the utilization of nanofluids in the
10
solar field in order to achieve higher system performance. An Organic Rankine Cycle
11
(ORC) driven by solar parabolic trough collectors (PTCs) is the examined system.
12
Four different nanoparticles are examined (Al2O3, CuO, TiO2 and Cu) in the base
13
fluid (Syltherm 800), as well as the pure thermal oil is examined as working fluid.
14
The examined ORC is a regenerative cycle and four organic fluids are also tested
15
(toluene, MDM, cyclohexane and n-pentane). In every combination between the
16
organic fluid in ORC and working fluid (nanofluid) in the solar field, an optimization
17
procedure is followed. The concentration of every nanoparticle and the pressure ratio
18
(pressure in the turbine inlet to critical pressure) are the optimization parameters.
19
According to the final results, toluene is the best organic fluid and CuO is the most
20
suitable nanoparticle. The combination of these two working fluids leads to
21
167.05kW electricity production and to 20.11% system efficiency with concentration
22
4.16%. The enhancement by the use of nanofluids is found up to 1.75% compared to
23
the respective case with pure thermal oil and this result indicates that the use of them
24
is able to improve the performance of solar driven ORCs. For the other nanoparticles
AC C
EP
TE D
M AN U
8
1
ACCEPTED MANUSCRIPT and Toluene in the ORC, Cu, Al2O3 and TiO2 lead to 166.18 kW, 165.72 kW and
26
165.60 kW electricity productions respectively with optimum concentrations 3.98%,
27
2.51% and 2.57% respectively.
28
Keywords
29
ORC, PTC, Nanofluid, Solar energy, Toluene, CuO/Syltherm
30
1. Introduction
31
Energy production and management are among the most discussed issues of the last
32
years due to many factors, as the fossil fuel depletion, the increasing energy need of
33
the humanity, the increasing price of electricity and the global warming problems [1-
34
2]. The exploitation of renewable and alternative energy sources as solar, geothermal,
35
wind, waste heat and biomass seems to be the sole sustainable solution. Among them,
36
solar energy presents many advantages, especially in countries with high solar
37
potential, and thus a lot of research has been focused on this energy source [3].
38
Solar collectors are the devices which capture the incident solar irradiation and
39
convert it into useful heat (solar thermal collectors), into electricity (photovoltaic
40
panels) or both into useful heat and electricity (hybrid collectors) [4]. However, at this
41
time, it is difficult for many solar-driven systems to compete for the conventional and
42
non-clean energy systems. In this direction, a lot of research has been focused on the
43
improvement of solar thermal collectors in order to increase their performance and to
44
make them the best candidate for thermal processes [5-6].
45
One of the most promising ways is the utilization of nanofluids are working fluids in
46
the solar collectors [7]. Nanofluids are working fluids which are created by adding
47
metallic nanoparticles inside a usual base fluid (water or thermal oil). The term of
48
nanofluids is introduced by Choi in 1995 for the first time [8]. The most usual
AC C
EP
TE D
M AN U
SC
RI PT
25
2
ACCEPTED MANUSCRIPT nanoparticles are the following: CuO, Cu, Fe, Al, Al2O3, TiO2, SiO2, and carbon
50
nanomaterial which have different thermal properties compared to the base fluid [9].
51
More specifically, these metallic particles present high thermal conductivity [8] and
52
density, two factors which can increase the thermal performance of the solar collector,
53
increasing the heat transfer coefficient in the working fluid (nanofluid) [10-11].
54
Up to this time, the majority of the literature studies are focused on the preparation of
55
nanofluids and of their thermal properties investigation (for instance in Refs [12-13]).
56
A smaller but increasing number of studies are focused on the performance of them in
57
the solar collector. Mwesigye et al. have been performed a great number of
58
simulations for nanofluid operation in parabolic trough collectors (PTCs) [14-16].
59
Mwesigye et al. [14] examined the use of Al2O3 inside the water for concentrations up
60
to 6% and they determined the optimum Reynolds numbers for various nanoparticle
61
concentrations. Mwesigye et al. [15] examined the use of Al2O3 in Syltherm 800 for
62
concentrations up to 8%. According to their results, thermal efficiency enhancement
63
up to 7.6% can be achieved. In the last study of these researchers [16], Cu
64
nanoparticle inside Therminol VP1 is investigated for concertation up to 6% and the
65
maximum thermal efficiency enhancement is found to be 12.6%. The use of Al2O3
66
inside thermal oils is examined also in Refs [17-19], as well as in Ref [20-21], with
67
encouraging results. Ghasemi et al. [22] compared the use of Al2O3 and CuO
68
nanoparticles in water base fluid by performing CFD simulations. According to their
69
results, both nanoparticles lead to thermal enhancement with CuO to be the best
70
candidate. Kasaeian et al. [23] examined the use of carbon nanotube/oil nanofluid in a
71
pilot PTC and they found significant enhancement. Moreover, other interesting
72
studies as the utilization of gas-phase nanofluids in transparent collectors are found in
73
the literature [24].
AC C
EP
TE D
M AN U
SC
RI PT
49
3
ACCEPTED MANUSCRIPT It is obvious that the examination of nanofluids in solar thermal collectors is a recent
75
topic of research and thus the examination of nanofluids in the thermal application has
76
not been extensively studied. There are an extremely small number of studies which
77
examine the nanofluids in greater solar-driven systems. More specifically, about one
78
or two studies in some applications can only be found in the literature. Kabeel and El-
79
Said [25] examined a hybrid solar desalination system with nanofluid (Al2O3/H2O) as
80
working fluid in a flat plate solar thermal collector. The experimental installation
81
included solar air heater, water solar collector, storage tank, humidifier and
82
dehumidifier. According to their results, the system performance is affected by the
83
nanoparticle concentration. Abu-Hamdeh and Almitani [26] investigated the use of
84
nanofluid in a solar driven liquid desiccant cooling system. They examined various
85
nanoparticles, as ZnO, Fe3O4, and Al2O3 and they found high thermal enhancements
86
up to 40% for various nanoparticle concentrations. Boyaghchi et al. [27] examined the
87
use of CuO nanofluid in the solar driven compression-absorption cascade refrigeration
88
system. They performed single and multi-objective optimization procedures in order
89
to determine the optimum values of the examined parameters, as well as the optimum
90
organic working fluid in the compression cycle. Finally, R134a is found to be the best
91
candidate. Their optimization parameters were the nanoparticle concentration, the
92
collector title angle, the collecting area, the intermediate pressure level and the mass
93
flow rate of the strong solution in the absorption chiller.
94
The next part of literature studies, which examines nanofluid in solar thermal
95
applications, is associated with the electricity production. Boyaghchi et al. [28]
96
examined a combined cooling, heating and power system which utilizes solar energy,
97
geothermal energy as well as there is auxiliary heat supply. The authors used
98
nanofluid as working fluid in solar thermal collectors and more specifically the used
AC C
EP
TE D
M AN U
SC
RI PT
74
4
ACCEPTED MANUSCRIPT CuO inside water. Their multi-objective optimization proved that R134a is the best
100
candidate as working fluid in the ORC. In another study with similar configuration,
101
the same nanofluid is examined [29]. The last two literature studies examine
102
electricity production system with Rankine cycles, driven by PTC, which operate with
103
nanofluids. Alashkar and Gadalla [30] examined three different nanoparticles (Cu,
104
Al2O3 and SWCNT) in Therminol and in Syltherm. They finally found that the Cu
105
nanoparticle to be the best candidate for both thermal oils. More specifically, the
106
optimum concentration of Cu is found to be 3% in the Therminol and 5% in Syltherm.
107
Toghyani et al. [31] examined the use of CuO, SiO2, TiO2 and Al2O3 nanoparticles in
108
thermal oil for feeding a water/steam Rankine cycle. They found that CuO is the best
109
nanoparticle and its optimum concentration is 4.28%.
110
As it is obvious from the previous literature review, the utilization of nanofluids in
111
solar thermal applications takes more and more attentions by the researchers. Up
112
today, there is a small number of studies which have examined the nanofluids in
113
thermodynamic cycles for electricity production. Thus, this study is devoted to the
114
detailed analysis and optimization of a solar driven ORC with parabolic trough
115
collectors. The innovative part of this study is the examination of various nanofluids
116
and organic fluids, as well as the optimization of all the possible combinations of the
117
previous working fluids. This strategy leads to a totally optimum system which
118
combines high thermal output in the solar field and high thermodynamic efficiency at
119
the ORC. The optimization of the energy systems is a very important issue in order to
120
design sustainable configurations [32]. The examined nanofluids are oil based, with
121
Syltherm 800 to be used. Al2O3, CuO, TiO2 and Cu are selected as the most usual and
122
efficient nanoparticles. Toluene, MDM, cyclohexane and n-pentane are the selected
123
organic fluids for the ORC. These fluids are selected because they have high critical
AC C
EP
TE D
M AN U
SC
RI PT
99
5
ACCEPTED MANUSCRIPT temperature and they can be optimally combined with PTC [33]. The optimization
125
parameters are the nanoparticle concentration and the pressure ratio which are directly
126
associated with the pressure in the turbine inlet of ORC. The analysis is performed
127
with EES (Engineering Equation Solver) in steady state condition in order to give the
128
emphasis in the thermal comparison of the different design cases.
129
2. Materials and Methods
130
2.1 The examined system
131
The examined system of this study is depicted in figure 1 with many details. This
132
system is examined with the commercial software EES (Engineering Equation
133
Solver). Parabolic trough collectors are selected for capturing the solar irradiation
134
(Qsol,t) and convert it into high-temperature heat (Qu,t). Eurotrough [34] module is
135
selected to be examined and 15 modules exist in the collector field. The working fluid
136
in this loop is nanofluid. The base fluid is Syltherm 800 [35] which is usual thermal
137
oil for operation from -40oC up to 400oC [35]. For different usual nanoparticles are
138
examined in this thermal oil: Al2O3, CuO, TiO2 and Cu. Moreover, the case of
139
operation with pure Syltherm 800 in the collector loop is examined in order to
140
perform suitable comparisons.
141
The storage tank includes pure Syltherm 800 and there is sensible heat storage. The
142
heat from the collector loop to the storage tank is transferred to a suitable heat
143
exchanger inside the storage tank. Hot thermal oil from the upper part of the storage
144
tank (Ts,in) goes to the heat recovery system (HRS) and it gives the demanded heat
145
input in the ORC (Qin). The cold thermal oil, after the HRS), has a temperature equal
146
to (Ts,out) and it is delivered in the down part of the storage tank.
AC C
EP
TE D
M AN U
SC
RI PT
124
6
ACCEPTED MANUSCRIPT The next step is the description of the regenerative ORC. Four different organic
148
working fluids are examined: toluene, MDM, cyclohexane and n-pentane. In the heat
149
recovery system, heat is transferred to the organic fluid and it becomes vapor. More
150
specifically, organic fluid in a liquid phase (state point 3) enters the heat recovery
151
system. Firstly in the economizer (ECO) part, it is converted to saturated liquid (state
152
point 34) and after in the evaporator (EVAP) heat exchange surfaces, it becomes
153
saturated vapor (state point 4). It is important to state that in the present study, the
154
system operates without superheating. The produced saturated steam goes to the
155
turbine where it expanded and electricity (Pel) is produced in the electrical generator
156
(G). The outlet of the turbine (state point 5) is a low-pressure superheated vapor of
157
high temperature. This steam is used in the recuperator in order the colder stream after
158
the condenser to be warmed up, a fact that leads to lower energy input demand in the
159
heat recovery system. More specifically, in the recuperator, the hot stream leaves this
160
device (state point 6) and enters to the condenser where heat is rejected to the
161
absorption heat pump in temperature level (Tcon). The saturated organic liquid (state
162
point 1) becomes subcooled liquid of high pressure after the pump (state point 2) with
163
electricity consumption (Wp) and this quantity is driven to the recuperator. In this
164
device, the enthalpy of the inlet stream (state point 1) increases and the warmer
165
stream (state point 3) enter the heat recovery system.
AC C
EP
TE D
M AN U
SC
RI PT
147
7
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
166
TE D
Figure 1. The examined solar ORC system
167
2.2 Working fluids
169
2.2.1 Nanofluids
170
In this study, four different nanoparticles are examined inside the base fluid (Syltherm
171
800). These nanoparticles are the following: Al2O3, CuO, TiO2 and Cu. The thermal
172
properties of these nanoparticles are given in table 1 [36]. It is important to state that
173
the thermal properties of Syltherm 800 have been taken by the EES libraries [35, 37].
AC C
174
EP
168
175
176
177
8
ACCEPTED MANUSCRIPT Table 1. Properties of the examined nanoparticles [36]
178
ρ (kg/m3) 3970 6320 4250 8933
k (W/mK) 40 77 8.95 401
cp (kJ/kgK) 0.765 0.532 0.686 0.385
RI PT
Particle Al2O3 CuO TiO2 Cu
The thermal properties of the nanofluids can be calculated according to equations 1 to
180
4, using the properties of the base fluid (bf) and of the nanoparticles (np). The density
181
of the mixture is given by equation 1 [38] and the specific heat capacity according to
182
equation 2 [39]:
183
ρ nf = ρ bf ⋅ (1 − φ ) + ρ np ⋅ φ ,
184
c p ,nf =
185
The thermal conductivity of the nanofluid is calculated according to the suggested
186
equation by Yu and Choi [40]:
M AN U
SC
179
TE D
ρ bf ⋅ (1 − φ ) ρ np ⋅ φ ⋅ c p ,bf + ⋅ c p ,np , ρ nf ρ nf
k np + 2 ⋅ k bf + 2 ⋅ (k np − k bf ) ⋅ (1 + β ) ⋅ φ
(1)
(2)
3
187
k nf = k bf ⋅
188
The parameter β is the ratio of the nanolayer thickness to the original particle radius
189
and usually, this parameter is taken equal to 0.1 for calculating the thermal
190
conductivity of the nanofluids [41].
191
The mixture viscosity can be calculated according to equation 4a and 4b [42-44]:
192
µ nf = µ bf ⋅ (1 + 2.5 ⋅ φ + 6.5 ⋅ φ 2 ) ,
193
µ nf =
k np + 2 ⋅ k bf − (k np − k bf ) ⋅ (1 + β ) ⋅ φ
,
(3)
AC C
EP
3
µ bf
(1 − φ )2.5
,
(4a)
(4b)
9
ACCEPTED MANUSCRIPT These equations are general and do not include the nanoparticle diameter and other
195
characteristics. It is essential to state that equation 4a is the Batchelor model and
196
equation 4b is the Brinkman model (ideal for spherical nanoparticles). These two
197
equations are evaluated and the give similar results (figure 3). Finally, equation 4a is
198
selected in this study.
199
The thermal properties of the examined nanofluids are presented in figures 2-3. In
200
these figures, the thermal conductivity, specific heat capacity, density and dynamic
201
viscosity are given for various temperatures (T) and nanoparticle concentrations (φ).
202
In every case, different depictions are selected in order to present the properties with
203
the clearest and simplest way.
204
Figure 2 shows the specific heat capacity (subfigures 2a and 2b), the density
205
(subfigures 2c and 2d) and the thermal conductivity (subfigures 2e and 2f). In
206
subfigures 2a, 2c and 2e, the four examined nanofluids with 3% concentration are
207
compared with the Syltherm 800 for temperatures from 25oC to 375oC. The other
208
subfigures (2b, 2d and 2f) compare the thermal properties of the four nanofluids for
209
different concentrations (from 0% to 6%) with a temperature equal to 300oC. The
210
followed presentation technique makes obvious the impact of temperature and of
211
concentration in these thermal properties. It is important to state that the
212
concentrations of 3%, as well as the temperature of 300oC, are representative values
213
for the examined cases in this work.
214
Subfigure 2a shows that the specific heat capacity of the nanofluids is lower than the
215
specific heat capacity of Syltherm 800. This result is explained by the low specific
216
heat capacity of the nanoparticles, as table 1 indicates. Al2O3 and TiO2 lead to similar
217
properties, while CuO and Cu lead to lower specific heat capacities respectively.
AC C
EP
TE D
M AN U
SC
RI PT
194
10
ACCEPTED MANUSCRIPT Subfigure 2b proves that higher concentration leads to lower specific heat capacity for
219
all the nanofluids. The density of nanofluids is greater than the Syltherm 800 case, as
220
subfigure 2c shows. Cu, CuO, TiO2 and Al2O3 create the order from the highest to the
221
lowest density values. Moreover, higher concentration leads to greater density,
222
according to subfigure 2d. Subfigures 2e and 2f prove that the thermal conductivity of
223
the examined nanofluids is similar and it is greater than the Syltherm 800 case. Also,
224
higher concentration leads to greater thermal conductivity, while higher temperature
225
leads to lower thermal conductivity.
AC C
EP
TE D
M AN U
SC
RI PT
218
226 227 228 229 230
Figure 2. a) Specific heat capacity for different temperatures b) Specific heat capacity for different concentrations c) Density for different temperatures d) Density for different concentrations e) Thermal conductivity for different temperatures f) Thermal conductivity for different concentrations
11
ACCEPTED MANUSCRIPT Figure 3 depicts the dynamic viscosity variation with temperature (subfigure 3a) and
232
concentration (subfigure 3b). According to equations 4a and 4b, the nanofluid
233
dynamic viscosity is depended on the base fluid viscosity and the concentration. Thus,
234
the present modeling leads to the same viscosity values among the examined
235
nanofluids. This is the reason for presenting only one general curve for nanofluids in
236
the subfigures of figure 3. Following the same methodology, as in figure 3, subfigure
237
3a compared the dynamic viscosity of the Syltherm 800 with the nanofluid case for
238
various temperatures. It is obvious that the nanofluid has higher viscosity with a small
239
difference with the Syltherm 800. Moreover, subfigure 3b proves that higher
240
concentration leads to higher nanofluid viscosity, a reasonable result according to
241
equation 4. Both models of equations 4a and 4b are depicted in figure 3. It is obvious
242
that the viscosity is approximately the same for both models and thus these models are
243
adopted as reliable. In this study, the model of equation 4a is selected to be used.
244
The results of figure 3 and 4 make clear the impact of the nanoparticles in the thermal
245
properties of the nanofluids. The basic conclusions from this mini-analysis are the
246
following: The nanofluids present higher density, thermal conductivity and viscosity,
247
compared to the base fluid. On the other hand, the specific heat capacity is lower. All
248
the thermal properties, except the specific heat capacity, have an increasing rate of the
249
increase in the nanoparticle concentration. Moreover, the density and the specific heat
250
capacity are depended on the kind of nanoparticle, with the thermal conductivity to be
251
influenced in an extremely small way and the viscosity to be the same among the
252
nanofluids.
AC C
EP
TE D
M AN U
SC
RI PT
231
12
RI PT
ACCEPTED MANUSCRIPT
SC
253
Figure 3. a) Dynamic viscosity between nanofluids and thermal oil for various
255
temperatures b) The impact of concentration in the dynamic viscosity for
256
temperature equal to 300oC
M AN U
254
2.2.2 Organic fluids
258
In the ORC, four different organic fluids are investigated. Toluene, cyclohexane,
259
MDM and n-pentane are the examined working fluid in this study. Table 2 includes
260
the examined working fluids and their basic properties as the critical temperature, the
261
critical pressure and the molecular weight, as well as the ASHRAE safety group of
262
every working fluid, ODP and GWP. Moreover, figure 4 shows the saturation lines of
263
the examined organic fluids in T-s figures. It is essential to state that these data are
264
taken from EES libraries [37], as well as the data of table 1 have taken from the same
265
libraries. All the examined working fluids are dry fluids and they are common in
266
similar studies with ORCs.
267
The most important parameter for all the working fluids is critical temperature
268
because it determines the maximum evaporating temperature level in the subcritical
269
cycle. Working fluids with critical temperature in the region close to 200-300oC are
270
selected to be examined because the heat sources temperature levels of this study are
AC C
EP
TE D
257
13
ACCEPTED MANUSCRIPT 271
close to this region. It is essential to state that the maximum examined heat source
272
temperatures in this study are close to 300oC – 350oC.
Toluene
MDM
Cyclohexane
n-pentane
350
RI PT
300
200
SC
T (oC)
250
150
M AN U
100 50 0 -1.5
-1.0
-0.5
0.0
0.5
1.0
1.5
s (kJ/kg K)
TE D
273
Figure 4. The saturation lines of the examined organic fluids in T-S diagram
275
Table 2. The examined organic fluids in the ORC [37, 45] Working
EP
274
Tcrit o
pcrit
MW
ODP
GWP ASHRAE group
( C)
(bar)
(kg/kmol)
Toluene
318.60
41.26
92.14
0
3
B3
MDM
290.90
14.15
236.5
0
n.a
A2
Cyclohexane
280.49
40.75
84.16
0
n.a
A3
n-pentane
196.55
33.70
72.15
0
5
A3
AC C
fluids
276
277 14
ACCEPTED MANUSCRIPT 2.3 Mathematical formulation
279
This subsection is devoted to describing the basic mathematical modeling if the
280
present system. The given equation concern energy balances, index definitions and
281
other useful modeling assumptions.
282
2.3.1 Solar field modeling
283
In this subsection, a detailed thermal modeling for the module Eurotrough is
284
presented. These equations can be combined together and finally, the thermal
285
efficiency of the solar collector can be calculated in every case.
286
Parabolic trough collectors are imaging concentrating collectors with high
287
concentration ratio and they exploit only the direct beam part of the incident solar
288
irradiation [20]. Thus, the available solar energy is calculated as the product of the
289
outer aperture (Aa) and the solar beam irradiation (Gb).
290
Qs = Aa ⋅ Gb ,
291
The outer absorber area is calculated according to equation 6, using the outer absorber
292
diameter (Dro) and the length (L) of the evacuated tube. The inner absorber area, as
293
well as the cover areas (inner and outer), can be calculated with similar formulas as
294
equation 6.
295
Aro = π ⋅ Dro ⋅ L ,
296
The useful energy that the heat transfer fluid gains are able to be calculated according
297
to the energy balance of its volume, as it is given in equation 7. It is important to state
298
that this quantity represents the useful heat of one module of the total system. The
299
specific heat capacity (cp) corresponds to the working fluid of the solar collector which
300
is nanofluid in the majority of the cases and pure thermal oil in only some cases.
(5)
AC C
EP
TE D
M AN U
SC
RI PT
278
(6)
15
ACCEPTED MANUSCRIPT 301
Qu = mcol ⋅ c p ⋅ (Tcol ,out − Tcol ,in ) ,
302
It is useful to state that the mass flow rate (m) is calculated as the product of the fluid
303
density (ρ) and the volumetric flow rate (V):
304
mcol = ρ ⋅ V / 3600 ,
305
The most important index for the evaluation of the solar collector is the thermal
306
efficiency (ηth). This parameter is calculated as the ratio of the useful energy to the
307
available solar energy:
308
η th =
309
If the solar collector modules are connected in parallel series, as in the present study,
310
then the thermal efficiency of one module is the same for the entire collector field [46].
311
So, the total useful energy of the collector field is calculated as:
312
Qu ,t = ηth ⋅ Qsol ,t ,
313
For “N” PTC modules, the total available solar irradiation Qsol,t is calculated as:
314
Qsol ,t = N ⋅ Qsol ,
315
The thermal losses of the absorber (Qloss), for one module, are radiation losses, as
316
equation 12 shows. It is important to state that in the evacuated tube collectors the heat
317
convection losses are neglected due to the vacuum between absorber and cover [47].
318
Qloss =
(7)
SC
RI PT
(8)
(9)
(10)
(11)
AC C
EP
TE D
M AN U
Qu , Qs
(
)
Aro ⋅ σ ⋅ Τr4 − Tc4 , 1 1 − ε c Aro + ⋅ εr ε c Aci
(12)
16
ACCEPTED MANUSCRIPT 319
The emissivity of the absorber (εr) is calculated as a function of its mean temperature
320
[48]:
321
ε r = 0.000327 ⋅ Tr − 0.065971,
322
In steady-state conditions, as in the present modeling, the thermal losses of the
323
absorber to the cover are equal to the thermal losses of the cover to the ambient. Cover
324
losses thermal energy due to radiation and to convection, as equation 14 shows [49]:
325
4 , Qloss = Aco ⋅ hout ⋅ (Tc − Tam ) + Aco ⋅ σ ⋅ ε c ⋅ Tc4 − Tsky
326
The sky temperature is calculated as [50]:
327
Tsky = 0.0553 ⋅ Tam ,
328
The heat convection coefficient between cover and ambient (hout) is estimated
329
according to equation 16 [49]:
330
0.58 hout = 4 ⋅ Vwind ⋅ Dco−0.42 ,
331
The wind velocity (Vwind) has a low impact on the results due to the evacuated tube. In
332
this study, this parameter is selected to be 1 m/s which leads approximately to hout = 10
333
W/m2K [49].
334
The energy balance on the absorber is a basic equation in the presented analysis
335
because this equation correlates the useful energy and the thermal losses, as equation
336
17 indicates. More specifically, this equation shows that the absorbed solar energy
337
(Qs· ηopt) is separated to useful heat and to thermal losses.
338
Qs ⋅ η opt = Qu + Qloss ,
SC
(14)
(15)
(16)
AC C
EP
TE D
1.5
)
M AN U
(
RI PT
(13)
(17)
17
ACCEPTED MANUSCRIPT 339
The optical efficiency (ηopt) is depended on the incident angle. The indecent angle
340
modifier K is used in order to calculate the optical efficiency of the collector for
341
various incidents angles [51].
342
ηopt (θ ) = ηopt (θ = 0ο )⋅ Κ (θ ) ,
343
In order to correlate the temperature level on the absorber and the fluid operational
344
temperature level, the heat transfer analysis inside the absorber tube has to be
345
examined. The next equation describes that heat transfer from the absorber to the
346
working fluid.
347
Qu = h ⋅ Ari ⋅ (Tr − T fm ) ,
348
The mean temperature of the working fluid is calculated as:
349
T fm =
350
A critical parameter of this modeling is the heat transfer coefficient (h) between
351
absorber tube and fluid. The tube geometry, the flow rate and the properties of the fluid
352
with the thermal conductivity play a significant role in the determination of the heat
353
transfer coefficient. The following equation shows the way that the heat transfer
354
coefficient can be calculated with the use of the dimensionless Nusselt number (Nu).
355
h=
356
Generally, the Nusselt number is determined with experimental analysis and there are
357
many literature equations about this number, for different operating conditions.
358
Usually, some other dimensionless numbers are used in these equations. These
359
numbers are presented below:
,
(19)
(20)
AC C
EP
TE D
2
M AN U
Tcol ,in + Tcol ,out
SC
RI PT
(18)
Nu ⋅ k , Dri
(21)
18
ACCEPTED MANUSCRIPT 360
Reynolds number (Re) in circular tubes is given below:
361
Re =
362
Prandtl number (Pr) is presented below:
363
Pr =
364
In this study, the Reynolds number is over 2300 and thus the flow is assumed to be
365
turbulent. More specifically, the Reynolds number is varied from 11,000 to 25,000
366
approximately. Different equations about the Nusselt number have been used for the
367
different working fluids in the PTC.
368
For the pure thermal oil case, the Dittus-Boelter equation for turbulent flow is used
369
[52]. This equation is valid for Reynolds numbers over 10,000 and Prandtl number
370
between 0.7 and 160.
371
Nu = 0.023 ⋅ Re 0.8 ⋅ Pr 0.4 ,
372
For Syltherm 800 and Al2O3, the equations 25, which is suggested by Pak and Cho
373
[53], is used. This equation is valid for Prandtl number between 6.5 and 12, for
374
Reynolds numbers between 10,000 and 100,000, and for concentrations up to 10%.
375
Nu = 0.021 ⋅ Re 0.8 ⋅ Pr 0.5 ,
376
For operation with Cu and CuO as nanoparticles inside the thermal oil, the equation of
377
Xuan and Li is applied [54]. The experiments of this Ref have been made for Reynolds
378
number between 10,000 and 25,000, while the concentration was up to 2%.
379
Nu = 0.059 ⋅ 1 + 7.68 ⋅ φ 0.6886 ⋅ (Re⋅ Pr )
4⋅m , π ⋅ Dri ⋅ µ
,
(23)
TE D
M AN U
SC
k
RI PT
µ ⋅ cp
(22)
AC C
EP
(24)
[
(25)
0.001
]⋅ Re
0.9238
⋅ Pr 0.4 ,
(26) 19
ACCEPTED MANUSCRIPT 380
For the case of TiO2, the equation of Duangthongsuk and Wongwises [41] is selected
381
to be used. The experiments of this Ref have been made for Reynolds number between
382
3,000 and 28,000, while the concentration was up to 2%.
383
Nu = 0.074 ⋅ Re 0.707 ⋅ Pr 0.385 ⋅ (100 ⋅ φ )
384
2.3.2 Storage tank and heat recovery system modeling
385
In the storage tank of the system, heat is stored in the thermal oil. This storage is
386
sensible and it is based on the increase on the thermal oil temperature increase. In the
387
present study, there is a heat exchanger in order the nanofluid not to be mixed with
388
the thermal oil. The heat exchanger is designed properly in order high heat amounts to
389
be transferred from the collector loop to the storage tank. The heat transfer from the
390
collector loop to the storage tank is assumed to be equal to the useful energy
391
production every time moment, a reasonable assumption for the steady state model.
392
This assumption practically leads to zero thermal storage on the collector loop,
393
something acceptable because of the low mass quantity in this loop tubes. So, the
394
following equation is used in this modeling:
395
Qu ,t = (UA)c − st ⋅
396
The total heat transfer coefficient (UA)c-st is taken equal to 12 kW/m2K in this study.
397
This parameter can be modified by designing greater or smaller heat exchanger, as
398
well as changing the shape of the heat exchanger area. It is also important to state that
399
the storage tank is assumed to have uniform temperature level equal to Tst. The
400
general energy balance on the storage tank is given below [46]:
401
Q stored = Qu ,t − Qloss − m s ⋅ c p s ⋅ (Ts ,in − Ts ,out ) ,
(27)
RI PT
,
TE D
M AN U
SC
0.074
EP
Tcol ,out − Tcol ,in
,
(28)
AC C
Tcol ,out − Tst ln Tcol ,in − Tst
(29) 20
ACCEPTED MANUSCRIPT 402
In steady state conditions, the stored energy (Qstored) is equal to zero. For the
403
definition of the thermal losses (Qloss), the following equation is used:
404
Qloss = U st ⋅ Ast ⋅ (Tst − Tam ) ,
405
The heat transfer coefficient (Ust) includes radiation, convection and conduction
406
losses and it is taken equal to 0.5 · 10-3 kW/m2K [55]; a value which corresponds to a
407
well-insulated storage tank. The outer area of the storage tank (Ast) can be calculated
408
according to Ref [55] for a cylindrical storage tank. The volume of the storage tank is
409
selected to be 10 m3. This parameter has a low impact on the system performance,
410
especially in steady state conditions as in this work.
411
On the other side of the system, heat is transferred to the ORC. Hot thermal oil from
412
the upper part of the storage tank with temperature (Ts,in) goes to the heat recovery
413
system and heat (Qin) is transferred in the ORC. The colder thermal oil, in the outlet
414
of the heat recovery system, has a temperature (Ts,out). Figure 5 depicts the general
415
heat exchange process inside the storage tank. The pinch point is observed at the start
416
of the evaporator and the temperature of the thermal oil at this point is calculated as:
417
Ts , m = Tsat + PP ,
418
At this study, the pinch point is taken equal to 20 oC, a typical value according to the
419
literature [33].
(31)
AC C
EP
TE D
M AN U
SC
RI PT
(30)
21
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
420
TE D
Figure 5. The heat transfer process in the heat recovery system
421
2.3.3 ORC modeling
423
The basic equations which describe the thermodynamic performance and operation of
424
the ORC are given in this subsection. These equations express mainly the energy
425
balances in the devices of the ORC.
426
The expansion in the turbine is modeled with the isentropic efficiency which is
427
defined as:
428
η is =
AC C
EP
422
h4 − h5 , h4 − h5is
(32)
22
ACCEPTED MANUSCRIPT The state point 5is has the same entropy with the state point 4 and its pressure is equal
430
to pl. This pressure level (pl) is the saturation pressure for heat rejection temperature
431
level (Tc).
432
The produced power (Pel) in the electrical generator is calculated according to
433
equation 33:
434
Pel = η mg ⋅ mo ⋅ (h4 − h5 ) ,
435
The power consumption (Wp) in the organic fluid heat pump is calculated as:
436
Wp =
437
The net power production (Pnet) of the ORC is the difference between the power
438
production in the electrical generator and the power consumption of the pump.
439
Pnet = Pel − W p ,
440
The high-pressure of the system is an important variable parameter in this study. The
441
pressure ratio (PR) is a dimensionless parameter which expresses the high-pressure.
442
This parameter is defined as the ratio of the high pressure to the critical pressure of
443
the fluid. The maximum value of this parameter is 0.9 for safety and stability reasons
444
[33].
SC
(33)
(34)
(35)
EP
TE D
M AN U
mo ⋅ ( ph − pl ) , ρ o ⋅η pum
AC C
445
RI PT
429
PR =
ph , p crit
(36)
446
The thermal efficiency (or thermodynamic efficiency) of the ORC (ηorc) is the ratio of
447
the net produced electricity to the heat input:
23
ACCEPTED MANUSCRIPT Pnet , Qin
448
η orc =
449
The heat input (Qin) can be calculated by the energy balance on the HRS by assuming
450
an ideal heat exchange:
451
Qin = mo ⋅ (h4 − h3 ) = m s ⋅ c ps ⋅ (Ts ,in − Ts ,out )
452
The system thermodynamic efficiency (ηsys) is calculated as:
453
η sys =
454
2.4 Methodology
455
In this study, the depicted system on figure 1 is examined and optimized. The basic
456
goal is to achieve maximum electricity production by keeping constant the collector
457
field area. Different working fluids are examined in the ORC and in the solar field
458
loop. More specifically, toluene, MDM, cyclohexane and n-pentane are the examined
459
organic fluids in the ORC cycle. In the solar field loop, four different nanofluids are
460
tested (Al2O3, CuO, TiO2 and Cu) inside Syltherm 800. Thus, four different
461
nanofluids, as well as pure Syltherm 800 are examined as working fluids in PTCs.
462
Totally twenty combinations of working fluids are investigated in this work.
463
At this point, it is essential to explain the way that this analysis is performed. The first
464
step (subsection 3.1) is the preliminary analysis of the system. In this analysis
465
(subsection 3.1.1), a simple strategy is followed in order to find how the examined
466
parameters (mainly nanoparticle concentration and pressure in the turbine inlet)
467
influence on the results. In this direction, the two main subsystems (solar field and
468
ORC) are examined separately. In the analysis of the PTC system, all the nanofluids
(38)
SC
,
RI PT
(37)
(39)
AC C
EP
TE D
M AN U
Pnet , Q s ,t
24
ACCEPTED MANUSCRIPT are examined for different concentrations ratios and for a typical inlet temperature in
470
the solar field (Tcol,in = 300oC). Moreover, the four organic fluids are compared for
471
different pressure ratios (PR). The best candidates from these sensitivity analyses are
472
selected to be examined deeper. In this analysis, different combinations of
473
nanoparticle concentration (φ) and pressure ratio (PR) are examined simultaneously
474
and the optimum combination of these parameters, which maximizes the system
475
electricity production (PTC and ORC) is found. In the next part of the preliminary
476
analysis, two different sensitivity analyses are performed (subsection 3.1.2). In the
477
first sensitivity analysis, the optimum nanofluid concentration is kept constant and the
478
pressure ratio is examined parametrically. In the second sensitivity analysis, the
479
pressure ratio is kept to its optimum value and the concentration is examined
480
parametrically. This strategy indicates how the parameters influence the system
481
performance and give reasons for the existence of an optimum solution.
482
The next step is the optimization of the total system with a more accurate way
483
(subsection 3.2). For all the working fluids combinations (20 combinations), an
484
optimization procedure is followed using the “Conjugate Directions Method” or
485
“Powell's method”. This method is supported by the utilized tool which is EES
486
(Engineering Equation Solver) by f-chart [37]. The relative convergence tolerance is
487
selected equal to 10-8 and a maximum number of iterations (function calls) is selected
488
to 104 (usually the solution was found after 200 to 300 iterations). A simple flow chart
489
for the optimization procedure is given in figure 6.
490
In this optimization procedure, the concentration of the nanoparticles is the first
491
optimization variable and it is varied from zero up to 6%. The second optimization
492
variable is the pressure ratio which varies from 10% up to 90%. For all the organic
493
fluids, 10% of pressure ratio is greater than the condensation temperature and so it
AC C
EP
TE D
M AN U
SC
RI PT
469
25
ACCEPTED MANUSCRIPT pressure ratio leads to acceptable values. Moreover, the upper limit is set to 90% in
495
order to keep a relatively safe distance to the critical point. It is important to note that
496
in the cases with pure Syltherm 800 in the collector loop, the nanoparticle
497
concentration is set to zero and only the pressure ratio is the optimization variable.
498
After performing twenty optimization procedures, the final optimum cases are
499
compared to each other in order to determine the global optimum solution. This global
500
optimum solution is a result of four different optimization parameters: nanoparticle
501
type, organic fluid, nanoparticle concentration and pressure ratio.
502
At the last part of this subsection, the basic parameters of the system simulation are
503
given in table 3. The majority of these parameters have typical values in order to
504
perform a representative study of the real systems. For the solar collector, the data
505
have been taken by Refs [48, 56] and the efficiencies have been taken from Ref [33].
506
Table 3. System constant parameters Parameters Values Ambient temperature (Tam) 25oC Solar beam irradiation (Gb) 0.8 kW/m2 Condenser temperature (Tcon) 40 oC Pinch point (PP) 20 oC Temperature difference in recuperator (∆Trc) 20 oC Electromechanical efficiency of the generator (ηmg) 98% Turbine isentropic efficiency (ηis) 85% Organic fluid pump efficiency (ηpum) 70% Thermal loss coefficient of the storage tank (Ust) 0.5·10-3 kW/m2K Storage tank volume (Vst) 10 m3 Heat exchanger effectiveness (UA)c-st 12 kW/m2K Volumetric flow rate on the collector module (V) 3 m3/h Number of modules (N) 15 Module collecting area (Aa) 69.2 m2 Collector module length (L) 12 m Incident angle modifier (K) ~1 Maximum optical efficiency [ηopt(θ=0ο)] 0.741 Absorber inner diameter (Dri) 0.066 m Absorber outer diameter (Dro) 0.070 m Cover inner diameter (Dci) 0.120 m Cover outer diameter (Dco) 0.125 m Cover emittance (εc) 0.90 Wind speed (Vwind) 1 m/s
AC C
EP
TE D
M AN U
SC
RI PT
494
26
ACCEPTED MANUSCRIPT It is essential to state that the solar irradiation has been kept constant during this
508
analysis. This strategy is followed in order to give the emphasis in the selection of the
509
proper working fluids and of the proper system characteristics. The selected value of
510
0.8 kW/m2 is a usual value in real applications and it is a reliable solution for sizing
511
the system in realistic conditions. It is useful to state that for the financial examination
512
of the system, the daily variation of the solar irradiation has to be examined, but this is
513
the objective of future studies on this system. It is important to state that the
514
developed model has been validated and used in other previous literature studies and
515
thus there is no reason for giving again theis validation evidence. More specifically,
516
the ORC modeling has been validated in Ref [33] and the parabolic trough collector
517
modeling has been validated in Refs [9, 57].
AC C
EP
TE D
M AN U
SC
RI PT
507
518 519
Figure 6. Flow chart of the optimization procedure
27
ACCEPTED MANUSCRIPT 3. Results
521
3.1 System performance with the parameter variation
522
3.1.1 Preliminary analysis
523
The objective of this subsection is to present the energetic behavior of the various
524
subsystems for different values of critical parameters. Firstly, the ORC is examined
525
and results for the thermal efficiency (ηorc) are presented in figure 7. In this figure,
526
different pressure ratios (Pratio) are applied for all the examined working fluids. In this
527
analysis, only the ORC is examined without the solar field. According to the results of
528
figure 7, higher pressure ratio leads to higher thermal efficiency. This result is based
529
on the simultaneous increase of the saturation temperature with the increase of the
530
high pressure or the pressure ratio. This fact makes the cycle to operate with a greater
531
high-temperature level and according to the ideal Carnot cycle, the thermal efficiency
532
is getting higher. The increase in the thermal efficiency is getting lower for greater
533
pressure ratios and the curves tend towards to horizontal. This fact is explained by the
534
decrease of (dTsat/dp) for higher pressures. The most efficient working fluid is toluene
535
with MDM, cyclohexane and n-pentane to follow respectively. The first three
536
working fluids lead to similar performance while the n-pentane leads to lower
537
performance. This result can be explained by the relatively lower critical temperature
538
(see table 2) of the n-pentane which creates a restriction in the system thermal
539
efficiency.
540
Figure 8 depicts the thermal efficiency of the solar collector (ηth) for the examined
541
nanofluids and for various concentrations. The zero concentration represents the
542
operation with pure Syltherm 800 and thus all the curves start from the same point.
543
These results concern operation with inlet temperature equal to 300oC. This
544
temperature level is representative for this study and thus it is selected. However, the
AC C
EP
TE D
M AN U
SC
RI PT
520
28
ACCEPTED MANUSCRIPT conclusions are the same for all the inlet temperature levels because in all cases, an
546
enhancement is achieved with the use of nanofluids. It is obvious that higher
547
nanoparticle concentration leads to higher thermal efficiency. The use of Al2O3 and
548
TiO2 leads to similar results and their curves are approximately the same. On the other
549
hand, the use of Cu and CuO presents similar behavior with Cu to leads to higher
550
thermal performance. Moreover, figure 8 shows that the use of CuO and Cu in the
551
thermal oil leads to a significant enhancement of the thermal performance, especially
552
for great concentrations.
MDM
0.35 0.30
0.20 0.15 0.10 0.05 0.00
554 555
0.3
EP
0.2
AC C
0.1
TE D
ηorc
0.25
553
Cyclohexane
n-pentane
M AN U
Toluene
SC
RI PT
545
0.4
0.5
0.6
0.7
0.8
0.9
Pressure Ratio - PR
Figure 7. Thermal efficiency of the ORC without the solar field for various working fluids and pressure ratios
29
ACCEPTED MANUSCRIPT 0.700
Syltherm 800 + Al₂O₃
Syltherm 800 + CuO
Syltherm 800 + TiO₂
Syltherm 800 + Cu
0.695
RI PT
ηth
0.690 0.685 0.680
0.00
0.01
0.02
0.03
φ
0.04
0.05
0.06
M AN U
556
SC
0.675
Figure 8. Thermal efficiency of nanofluids for various concentrations with the
558
inlet temperature equal to 300oC
559
Figures 7 and 8 proved that toluene and Syltherm 800 with Cu are the best candidates
560
for utilization in the examined system. Thus, the operation with these fluids is
561
examined with more details in this subsection. Figure 9 shows the electricity
562
production for various concentrations and for six pressure ratios (from 0.4 to 0.9). In
563
every curve of this figure, the maximum point is depicted. According to the results,
564
the global maximum is observed for pressure ratio equal to 0.8 and concentration
565
2.50%, with the electricity production to be 166.19 kW. The next case is for pressure
566
ratio 0.7 and concentration 2.75% where the electricity production is 166.18 kW.
567
These solutions are similar and maybe equivalent. The next choice is for pressure
568
ratio 0.6 and 3% concentration with 165.06 kW. An interesting observation is that the
569
optimum concentration is getting lower for higher pressure ratios and it is ranged
570
from 2.25% (for PR=0.9) to 3.75% (for PR=0.4). Figure 10 is a three-dimensional
571
depiction of figure 9 and it is presented for giving a better image of the optimization
AC C
EP
TE D
557
30
ACCEPTED MANUSCRIPT need. In other words, the green-top area in the electrical output surface is the optimum
573
area. All the optimum solutions are located in the projection of this area to the
574
horizontal surface (PR - φ). This observation proves the need of a deeper optimization
575
procedure in order to determine in every case an optimum couple of pressure ratio and
576
nanoparticle concentration.
577
The results of figures 9 and 10 prove that there is an internal optimum solution which
578
maximizes the electricity production. However, this result is maybe strange because
579
the separate analysis in figures 7 and 8 does not indicate this. More specifically, figure
580
7 shows that higher pressure ratio leads to greater ORC efficiency and figure 8 shows
581
that higher nanoparticle concentration leads to higher solar thermal collector
582
efficiency. Figures 9 and 10 prove that the combination of the solar system and the
583
ORC operates optimally for intermediate values of concentration and pressure ratio.
584
This result is very interesting and it will be examined in subsection 3.1.2 in details.
585
Moreover, the system efficiency (ηsys) is illustrated in figure 11. This figure proves
586
that the system efficiency surface has a similar shape with the one of the electricity
587
production (figure 10). The maximum efficiency is 20.00%, while the lowest system
588
efficiency, for the depicted cases, is 18.85%.
AC C
EP
TE D
M AN U
SC
RI PT
572
31
ACCEPTED MANUSCRIPT PR = 0.8 PR = 0.5
PR = 0.7 PR = 0.4
RI PT
167 166 165 164 163 162 161 160 159 158 157 156 0.00
0.01
0.02
0.03
φ
0.05
0.06
M AN U
589
0.04
SC
Pel (kW)
PR = 0.9 PR = 0.6
Figure 9. Electricity production for operation with toluene and Cu nanofluid.
591
The red points represent the maximum points of the examined curves.
592
AC C
EP
TE D
590
593
Figure 10. Three-dimensional depiction for the optimization procedure of
594
electricity output. These results correspond the case with toluene and Cu
595
nanofluid.
32
0.1975-0.2000 0.1900-0.1925 0.1825-0.1850
0.1950-0.1975 0.1875-0.1900 0.1800-0.1825
0.2025 0.2000 0.1975 0.1950 0.1925 0.1900 0.1875 0.1850 0.1825 0.1800
RI PT
0.2000-0.2025 0.1925-0.1950 0.1850-0.1875
ηsys
ACCEPTED MANUSCRIPT
0.0%
1.0%
2.0% 3.0% 4.0% 0,5
PR
5.0% 6.0%
φ
SC
596
0,7
M AN U
0,9
597
Figure 11. System efficiency for various concentrations and pressure ratios.
598
These results correspond the case with toluene and Cu nanofluid. 3.1.2 Parametric analysis
600
In order to perform a deeper analysis, two cases will be examined with more details.
601
The first is for constant nanoparticle concentration and variable pressure ratio, while
602
the second is for constant pressure ratio and different nanoparticle concertation. This
603
method will try to approximate the maximum point with two lines, one parallel to the
604
nanoparticle axis (see the black line in figure 10) and the other parallel to the pressure
605
ratio axis (see the red line in figure 10).
606
Figures 12 to 14 are devoted to the first case with constant nanoparticle concentration
607
(2.5%). Figure 12 shows that the saturation temperature increases with the increase of
608
the pressure ratio. This result makes all the system temperature levels to increase.
609
More specifically, the thermal oil temperature in the heat recovery system is getting
610
greater with the increase of the saturation temperature. The thermal oil outlet
611
temperature (Ts,out) has similar values with the saturation temperature, as the Q-T
AC C
EP
TE D
599
33
ACCEPTED MANUSCRIPT diagram of figure 5 also indicates. The inlet temperature in the HRS (Ts,in) is higher
613
than (Ts,out) in order to give the demanded heat to the ORC. The temperature levels of
614
the nanofluid in the collector loop are higher that the thermal oil temperature levels in
615
the thermal oil loop (storage tank and HRS). The inlet temperature in the collector
616
field (Tcol,in) is a bit higher than the storage tank temperature (which equal to Tl,in in
617
the present modeling), while the collector field outlet temperature (Tcol,out) is the
618
highest temperature in the system.
619
The temperature level variation, which is depicted in figure 12, influence on the
620
various subsystem efficiencies (ORC and PTC), as well as in the total system
621
efficiency (ηsys). All these indexes are given in figure 13 for various pressure ratios.
622
Higher pressure ratio leads to higher temperature level in all the points of the system,
623
a fact that makes the subsystem to have different behavior. The higher temperature in
624
the solar collector makes its thermal efficiency to get lower. On the other hand, higher
625
temperature in the ORC and especially in the turbine inlet (or higher saturation
626
temperature) makes the thermodynamic efficiency to be higher. The system
627
performance is strongly influenced by these two parameters and thus the system
628
efficiency presents maximum point for an intermediate pressure ratio. This is a very
629
interesting result and it proves the need for a detailed optimization procedure of the
630
cycle. Figure 13 shows that the thermal efficiency on PTC is ranged from 65.59% to
631
71.43%, the ORC efficiency from 20.97% 30.52% and the system efficiency from
632
14.90% to 20.00%. The system efficiency is maximized for pressure ratio equal to
633
0.8, and for this pressure ration the PTC efficiency is 66.13%, the ORC efficiency is
634
30.52% and the system efficiency 20.00%.
AC C
EP
TE D
M AN U
SC
RI PT
612
34
ACCEPTED MANUSCRIPT Tcol,out
Tcol,in
Ts,in
Ts,out
Tsat
400
300
RI PT
Temperature (oC)
350
250 200 150 0.2
0.3
0.4
0.5
0.6
0.7
SC
0.1
0.8
0.9
Pressure ratio - PR
M AN U
635
Figure 12. Temperature levels in the system for various pressure ratios when the
637
Cu concentration is 2.5% and toluene is the organic working fluid
638
The next step is to investigate the energy flow inside the system. Figure 14 depicts the
639
basic heat and electricity streams in the system for various pressure ratios. More
640
specifically, the electricity production (Pel), the heat input in the ORC (Qin) and the
641
useful energy production in the collector field (Qu,t) are given for various pressure
642
ratios in this figure. The heat input in the ORC is lower than the useful energy
643
production in the collector field (Qin < Qu,t) due to the thermal losses in the storage
644
tank. The electricity production is a small fraction of the heat input in the ORC
645
because of the relatively small thermodynamic efficiency of the cycle (ηorc).
646
The electricity production curve presents maximum for pressure ratio close to 0.8.
647
This result is interesting and it can be explained by the results of figure 14. For this
648
pressure ratio, the electricity production is 166.19 kW, the heat input in the system is
649
544.4 kW and the useful heat production in the collector field is 549.4 kW.
AC C
EP
TE D
636
35
ACCEPTED MANUSCRIPT System
ORC
PTC
0.8 0.7
0.5 0.4
RI PT
Efficiency
0.6
0.3 0.2 0.1 0.0 0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
SC
0.1
Pressure ratio - PR
M AN U
650 651
Figure 13. Efficiencies (ORC, PTC and system) for various pressure ratios when
652
the Cu concentration is 2.5% and toluene is the organic working fluid
Pel
TE D
170 165 160 155 150 145 140 135 130 125 120
580 570 560
EP
Energy (kW)
590
Qu,t
550
AC C
540 530
0.1
653
0.2
0.3
0.4
0.5
0.6
0.7
0.8
Electricity (kW)
Qin 600
0.9
Pressure ratio - PR
654
Figure 14. Electricity production, useful energy production in the collector field
655
and heat input in the ORC for various pressure ratios (Cu concentration 2.5%
656
and toluene)
36
ACCEPTED MANUSCRIPT The next step in this analysis is to examine the cases of constant pressure ratio equal
658
to 0.8 and the nanoparticle concentration to be a variable parameter. Figures 15 and
659
16 include the respective results for this case with toluene and Cu nanoparticles.
660
Figure 15 gives the various temperature levels in the system and the most variable
661
properties of the nanofluid with the different concentrations (specific heat capacity
662
and density). These thermal properties are calculated for the mean nanofluid
663
temperature in the solar collector. It is obvious that the temperature difference in the
664
nanofluid circuit (Tcol,out - Tcol,in) is getting lower with the increase in concentration
665
(blue color curves). The main reason for this result is the variation of the thermal
666
properties. Higher density is achieved with the increase in the concertation and this
667
fact makes the mass flow rate to be higher. This situation makes the temperature
668
difference to be lower for higher φ. However, the simultaneous decrease in the
669
specific heat capacity comes to counterbalance this situation and the phenomenon is
670
not very intense. The temperature difference in the thermal oil circuit (black color
671
curves) has also decreasing rate. This result is explained by the increase in the thermal
672
oil specific capacity and the decrease in the heat transfer quantity in high
673
concentrations, as figure 15 shows.
674
Figure 16 proves that there is an optimum concentration where the electricity
675
production is maximized. For the present case, this concentration is 2.5% and the
676
electricity production is 166.19 kW. The heat input in the system is 544.4 kW and the
677
useful heat production is 549.4 kW. These numbers are the same as it is presented in
678
figure 14. This result is reasonable because the optimum point of figure 14 is the same
679
to the optimum point of figure 16. This optimum point is the global optimum point of
680
figure 10 and it is the cross-section of the read and the black curves in figure 10.
AC C
EP
TE D
M AN U
SC
RI PT
657
37
ACCEPTED MANUSCRIPT It would be valuable to state that the existence of the optimum concentration is an
682
important result which has to be explained more. The decreasing rate of temperature
683
difference (Tcol,out - Tcol,in) in figure 15, leads to lower heat input (Qin) in the storage
684
tank. On the other hand, higher concentration leads to greater thermal efficiency
685
(figure 8). These reverse factors lead to an optimum concentration which leads to
686
maximum system performance.
Tcol,in
Ts,in
Ts,out
SC
355 345
M AN U
Temperature (oC)
ρ
335 325 315 305 295 275 0.00
TE D
285 0.01
0.02
0.03
cp 2500 2000 1500 1000 500
ρ (kg/m3) , cp (J/kg K)
Tcol,out 375 365
RI PT
681
0 0.04
0.05
0.06
concentration - φ
687
Figure 15. Temperature levels in the system, nanofluid density and nanofluid
689
specific heat capacity for various concentrations of Cu nanoparticles (PR=0.8)
AC C
EP
688
38
ACCEPTED MANUSCRIPT Qin
Qu,t
Pel
552
166.5
550
544
165
542
RI PT
Energy (kW)
165.5
546
164.5
540
Electricity (kW)
166
548
164
538 536
163.5
0.01
0.02
0.03
0.04
concentration - φ
690
0.05
0.06
SC
0.00
Figure 16. Electricity production, useful energy production in the collector field
692
and heat input in the ORC for various concentrations of Cu (PR=0.8)
693
The results of subsection 3.1.2 showed that for every pressure ratio, there is an
694
optimum concentration and for every concentration, there is an optimum pressure
695
ratio which maximizes the electricity production. These results are taken for the case
696
with Cu nanoparticle and toluene as organic fluid.
697
3.2 Optimization results
698
The results of the previous analysis proved that there is need of systematic
699
optimization. In this section, the optimum pressure ratio and concentration are
700
determined for all the combinations between nanoparticles and organic working
701
fluids. A more detailed methodology is followed, by using the “Conjugate Directions
702
Method” which is supported by EES. More specifically, for every combination of
703
organic fluid and fluid in the PTC circuit, an optimization procedure is followed. In
704
this procedure, the nanofluid concentration and the pressure ratio are the optimized
705
parameters. For the case of thermal oil in the PTC circuit, the concentration is kept
AC C
EP
TE D
M AN U
691
39
ACCEPTED MANUSCRIPT equal to zero in the optimization. So, twenty different combinations are examined and
707
the results are given below in figures 17 to 20 and in table 4.
708
Figure 17 illustrates the electricity production for the all the optimized cases. It is
709
obvious that toluene is the best working organic fluid in all the cases. MDM comes
710
second with cyclohexane and n-pentane to follow respectively. It is important to state
711
that n-pentane is the less suitable working fluid for this system because of its lower
712
critical temperature which makes restriction in the maximum work production from
713
the thermodynamic cycle. Among the nanofluids, CuO is the best candidate and it
714
leads to 167.05 kW electricity production combined with toluene. The next candidate
715
is Cu with 166.32 kW, while Al2O3 and TiO2 follow with 165.72 kW and 165060 kW
716
respectively. The case with pure thermal oil leads to the lowest electricity production
717
with 164.18 kW. This consequence (CuO, Cu, Al2O3, TiO2, thermal oil) in the
718
performance is the same for all the organic fluids and this is an important result.
719
Moreover, it is useful to comment about the exact electricity productions for all the
720
organic fluids with the best nanoparticle (CuO). As it has been said, the electricity
721
production with toluene is 167.05 kW. MDM, cyclohexane and n-pentane lead to
722
164.86 kW, 160.01 kW and 117.02 kW respectively.
723
It is obvious that the impact of the organic fluid selection in higher on the electricity
724
production than the impact of the proper nanofluid selection. This result indicates that
725
the utilization of the best organic working fluid is the major step in every optimization
726
procedure. On the other hand, the use of nanofluid leads always to higher
727
performance, compared to the pure thermal oil case. This result indicates the use of
728
nanoparticles inside the base fluid in order to achieve higher performance in all the
729
cases. Figure 18 shows the system efficiency results for all the optimized cases. The
730
results of this figure are similar to the results of figure 17 but they are expressed in
AC C
EP
TE D
M AN U
SC
RI PT
706
40
ACCEPTED MANUSCRIPT efficiency terms. More specifically, again toluene is the best organic working fluid
732
and CuO is the best nanoparticle. The system efficiency with toluene is 20.11% with
733
CuO, 20.02% with Cu, 19.95% with Al2O3 and 19.93% with TiO2. The system
734
efficiency with pure thermal oil is 19.76% and it is the lower than all the cases with
735
nanofluids.
Toluene 180
MDM
167.05
165.72
Cyclohexane 165.60
n-pentane
166.32
140
100 80 60 40 20 0
Al₂O₃
TiO₂
TE D
CuO
M AN U
Pel (kW)
120
737
164.18
SC
160
736
RI PT
731
Cu
Thermal oil
Figure 17. Optimum electricity production for all the examined cases
0.25
0.1995
0.2011
Cyclohexane
0.1993
n-pentane
0.2002
0.1976
AC C
0.20
MDM
EP
Toluene
ηsys
0.15 0.10 0.05 0.00
738 739
Al₂O₃
CuO
TiO₂
Cu
Thermal oil
Figure 18. Optimum system efficiency for all the examined cases 41
ACCEPTED MANUSCRIPT Figures 19 and 20 exhibit the optimum values of the variable parameters. Figure 19
741
gives the optimum nanoparticle concentration for all the examined cases. It is
742
important to state that for the cases with thermal oil, the concentration is set equal to
743
zero. Figure 20 gives the optimum values of the pressure ratio for all the cases. The
744
optimum concentration, in figure 19, presents significant variations in the examined
745
range (0% - 6%). Generally, the n-pentane performs better with the highest possible
746
concentrations close to the upper limit (~6%). Moreover, the Al2O3 inside the
747
cyclohexane obtains it upper limit value for the optimum operation. In the other cases,
748
concentrations from 2.5% to 4.35% are found as optimum. Toluene is the working
749
fluid which needs the lowest concentrations, with MDM, cyclohexane and n-pentane
750
to need higher concentrations respectively. Al2O3 and CuO have to be used in higher
751
concentrations, compared to TiO2 and Cu, as the results proved. For example, toluene
752
needs 4.16% Al2O3 and 3.98% CuO for optimum operation, while 2.57% TiO2 and
753
2.51% Cu. Taking into account all the above, as well as the ORC efficiency results of
754
figure 7, the most efficient working fluids (toluene and MDM), need relatively lower
755
nanofluid concentration in order to achieve maximum system performance. On the
756
other hand, the less efficient working fluids, n-pentane mainly and cyclohexane in
757
some cases, need high concentrations in the nanofluid in order the higher thermal
758
efficiency in the PTC to counterbalance the lower ORC efficiency.
759
Figure 20 depicts the optimum pressure ratios for all the examined cases. It is useful
760
to state again that the maximum value of this parameter is set to 90% in order not to
761
go extremely close to the critical point. Generally, values between 75% and 90% are
762
the optimum, a fact that proves the need for high pressure in the turbine inlet. It is also
763
obvious that the kind of nanofluid (or the pure thermal oil case) does not play a
764
significant role on the optimum pressure ratio. So, for toluene, a pressure level close
AC C
EP
TE D
M AN U
SC
RI PT
740
42
ACCEPTED MANUSCRIPT to 75% is optimum for all the cases. MDM needs to operate at the highest possible
766
pressure level (PR=90%), while cyclohexane and n-pentane need pressure ratio close
767
to 87%. Practically, the exact curvature of the saturation line is the determining factor
768
for the optimum pressure level in every fluid. As a final result, the pressure ratio has
769
to be selected close to highest possible values except for the case of toluene.
770
However, it is important to state that toluene has the highest critical temperature
771
compared to the other examined organic fluids and this also makes the optimum
772
pressure ratio to bit relatively lower, in order to achieve the suitable (optimum)
773
saturation temperature inside the HRS.
M AN U
Toluene
MDM
0.06 0.05
0.01 0.00
775
n-pentane
Al₂O₃
CuO
TiO₂
Cu
Thermal oil
AC C
774
Cyclohexane
TE D
0.02
EP
φopt
0.04 0.03
SC
RI PT
765
Figure 19. Optimum concentration for all the examined cases
43
ACCEPTED MANUSCRIPT Toluene
MDM
Cyclohexane
n-pentane
0.90 0.85
RI PT
0.75 0.70 0.65
Al₂O₃
CuO
TiO₂
Cu
Thermal oil
M AN U
776
SC
PRopt
0.80
Figure 20. Optimum pressure ratio for all the examined cases
778
Table 4 summarizes the presented results of figures 17 to 20, as well as includes data
779
about the ORC efficiency, the PTC efficiency and the optimum saturation
780
temperatures. It is useful to be commented that the thermal efficiency in the solar field
781
is close to 67%, while the ORC efficiency is close to 28-30% for the three more
782
efficient fluids (toluene, MDM, cyclohexane) and close to 20% about n-pentane. The
783
system efficiency is approximately the product of the previous efficiencies and it is
784
close to 20% for the three more efficient working fluids and close to 14% for n-
785
pentane. The saturation temperature which leads to the optimum operation is variable
786
among the examined working fluid. It is close to 296oC for toluene, 284oC for MDM,
787
269oC for toluene and 187oC for n-pentane. The critical temperature is one parameter
788
which influences on the previous optimum temperature level by creating restrictions
789
in the highest possible saturation temperature (remember the constraint of 90% higher
790
pressure ratio).
AC C
EP
TE D
777
44
ACCEPTED MANUSCRIPT It is valuable to comment about the impact of nanoparticle kind on the final results.
792
Comparing the optimized cases with the same organic fluid, it is obvious that there is
793
a small difference in the system efficiency among the different nanofluids. These
794
results can be explained by two factors. The first one is that the final results are
795
optimized and for every nanofluid, the optimum parameters have been obtained. Thus,
796
this result indicates that all the examined nanoparticles can lead to system
797
performance enhancement. The second factor is based on the high PTC thermal
798
efficiency which gives a small margin of thermal improvement by the use of
799
nanoparticles.
800
Moreover, it is essential to state in the optimum nanoparticles concentrations. Table 4
801
shows that different concentrations are optimum among the examined cases. This
802
result indicates the need of optimization for determining the most suitable
803
concentration in every case. The reason for the existence of optimum concentration
804
has been explained in subsection 3.1.2 and especially with figures 15 and 16. More
805
specifically, for some nanofluids, higher nanoparticle concentration leads to a
806
different temperature in the collector field and in the storage tank and this fact makes
807
the system performance to be variable.
808
Figure 21 illustrates the temperature-specific entropy diagram of the process. This
809
figure depicts the optimum case with toluene and CuO nanofluid. The concentration is
810
3.98% and the pressure ratio 75.51% in order to give the best case. It is obvious that
811
the ORC covers a great part of the saturation area, a fact that proves great temperature
812
difference between the turbine inlet and the condensation temperature. Moreover, the
813
heat transfer from the nanofluid to the thermal oil and from the thermal oil to the
814
organic fluid is also given. It is important to state that the heat transfer from the
815
nanofluid to the thermal oil is achieved with the storage tank, while the heat transfer
AC C
EP
TE D
M AN U
SC
RI PT
791
45
ACCEPTED MANUSCRIPT 816
from the thermal oil to the nanofluid is done in the heat recovery system. It is also
817
useful to state that the state point 7 is inside the condenser and this is the saturation
818
vapor in the low pressure.
ηorc (%) 30.42 30.42 30.42 30.42 30.43 29.63 29.63 29.63 29.63 29.62 28.42 28.43 28.43 28.43 28.43 19.80 19.80 19.80 19.80 19.79
φ (%) 4.16 3.98 2.57 2.51 0.00 4.35 4.20 2.70 2.66 0.00 6.00 4.75 4.16 3.04 0.00 6.00 6.00 6.00 5.88 0.00
TE D
PR (%) 75.25 75.51 75.48 75.22 75.79 90.00 90.00 90.00 90.00 90.00 86.68 86.83 86.78 86.71 87.13 86.78 86.80 86.75 86.64 86.87
SC
ηth (%) 66.17 66.68 66.10 66.41 65.54 67.05 67.53 67.00 67.26 66.47 67.82 68.28 67.76 68.03 67.18 71.17 71.51 71.15 71.39 70.51
M AN U
ηsys (%) 19.95 20.11 19.93 20.02 19.76 19.70 19.84 19.68 19.76 19.52 19.12 19.26 19.11 19.19 18.94 14.01 14.08 14.01 14.06 13.88
Tsat (oC) 295.6 295.9 295.8 295.5 296.1 283.9 283.9 283.9 283.9 283.9 269.1 269.2 269.2 269.1 269.5 186.7 186.8 186.7 186.6 186.8
AC C
820
Al2O3 CuO TiO2 Cu Thermal oil Al2O3 CuO TiO2 Cu Thermal oil Al2O3 CuO TiO2 Cu Thermal oil Al2O3 CuO TiO2 Cu Thermal oil
Pel (kW) 165.72 167.05 165.60 166.18 164.18 163.66 164.86 163.55 164.19 162.21 158.90 160.01 158.77 159.41 157.41 116.45 117.02 116.41 116.82 115.35
EP
n-pentane
Cyclohexane
MDM
Toluene
Working fluids
RI PT
Table 4. Optimum results for all the examined cases
819
46
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
821 822 823
Figure 21. Temperature – specific entropy diagram for the optimum case with toluene and CuO nanofluid (φ = 3.98% and PR = 75.51%) 3.3 Discussion
825
This subsection is devoted to summarizing the results and to discussing them on a
826
deeper basis. This work is an optimization of a solar driven ORC by using nanofluids
827
inside the PTCs. The main scope of this work is to determine the performance
828
enhancement of the system by using nanofluids, compared to the conventional
829
thermal oil case. Figure 22 gives the enhancement for all the examined cases with
830
nanofluids compared to the respective with pure thermal oil and of course the same
831
organic working fluid. The highest enhancement is found with CuO and they are close
832
to 1.7%. Cu case follows with 1.25%, while Al2O3 leads to 0.95% enhancement and
833
TiO2 to 0.85% enhancement. This enhancement is in the electrical production but it is
834
the same for the system performance. It is essential to state that the comparison is
835
performed with the optimized thermal oil case for the same organic fluid in every
AC C
EP
TE D
824
47
ACCEPTED MANUSCRIPT case. So, the results are compared to also optimized cases and the shown enhancement
837
is associated only with the nanofluid utilization.
838
The global maximum enhancement is observed for toluene with CuO nanofluid and it
839
is 1.75%. The next is with cyclohexane and CuO (1.65%) and the third with MDM
840
and CuO (1.63%). Generally, it is obvious that the thermal enhancement is not so
841
great but it is important for improving the performance of the systems. If financial and
842
environmental parameters are taken into account, this enhancement is able to make
843
the investment more sustainable in the total project life horizon. However, the higher
844
cost of working fluid with the nanoparticles, as well as the preparation cost and the
845
operating - maintenance costs have to be taken into account for the final decision of
846
the nanofluid utilization. Moreover, it would be important to state that the equations
847
about the thermal properties of nanofluids (equations 1-4) and the equations about the
848
Nusselt number (equations 25-27) have taken from literature studies. This strategy
849
leads to reliable results but an extra experimental for them would be important and it
850
can be performed in the future.
TE D
M AN U
SC
RI PT
836
1.6% 1.4%
1.75%
1.2% 1.0%
Cyclohexane
n-pentane
1.65% 1.63% 1.45% 1.27% 1.30% 1.27% 1.22%
AC C
Performance enhancement
1.8%
MDM
EP
Toluene
0.95% 0.94% 0.95% 0.89%
0.86% 0.86% 0.92% 0.83%
0.8% 0.6% 0.4% 0.2% 0.0%
851 852
Al₂O₃
CuO
TiO₂
Cu
Figure 22. Performance enhancement with the use of nanofluids 48
ACCEPTED MANUSCRIPT Another interesting result of this study is the results about the best nanoparticle.
854
According to the preliminary study of the solar collector (see figure 8), the best
855
nanoparticle is Cu. However, the analysis of the total system proves that CuO is the
856
best candidate (see for example figure 21). This result is very interesting and proves
857
that a combined analysis (with solar and ORC together) has to be made in order to
858
determine the optimum nanofluid for the examined application. As a final conclusion,
859
the utilization of CuO is the most suitable in the solar driven system because the final
860
optimization results indicated this as the best candidate. However, the performance
861
different between CuO and Cu cases are relatively small.
862
In the last part of the discussion section, it is important to discuss about the
863
assumptions of this study. The adoption of equations 26 and 27 for concentrations up
864
to 6% is an important assumption of this study because the respective experimental
865
studies have been performed for concentrations up to 2%. Moreover, the selection of
866
viscosity models which take into account only the concentration (φ) is a questionable
867
issue. Thus, an extra analysis is given below. The following model of Corcione [58] is
868
selected to be used as a reliable model which takes into account the nanoparticle
869
diameter (dp), as well as the equivalent diameter of the base fluid molecule (df). The
870
nanofluid viscosity is given as [58]:
871
µ nf =
AC C
EP
TE D
M AN U
SC
RI PT
853
µ bf
dp 1 − 34.87 ⋅ d f
,
− 0.3
(40)
⋅ φ 1.03
872
The equivalent diameter of the base fluid molecule is given as [58]:
873
6 ⋅ MW d f = 0.1 ⋅ N ⋅π ⋅ ρ bf a
1
3 ,
(41)
49
ACCEPTED MANUSCRIPT For Syltherm 800, the molecular weight (MW) is 317 kg/kmol, the density at 293 K
875
(ρbf) is 935.3 kg/m3, while the Avogadro number (Na) is equal to 6.022 · 1026 kmol-1.
876
The nanoparticle diameter (dp) has taken equal to 100 nm or 10-7 m for all the
877
nanoparticles. This diameter selection is the same as in Refs [53-54] and it is a
878
representative value for performing a simple comparative analysis. Equation 40 is
879
valid for concentrations up to 7%, for nanoparticle diameters from 25 nm to 200 nm
880
and for temperatures from 20oC to 60oC. Figure 23 give the comparison between the
881
calculation of viscosity with equation 4a (used in this paper) and equation 40. It can
882
be shown that in low concertation up to 3%, the curves are relatively close. For higher
883
concentration, their difference is getting higher with deviation up to 20%
884
approximately. Figure 24 shows the thermal efficiency of the collector using the two
885
viscosity models when the inlet temperature is equal to 300oC. This temperature level
886
is characteristic of the examined system and thus it is selected as a representative
887
value for this analysis. It is obvious that the thermal efficiency is about the same with
888
the two viscosity models with maximum deviation about 0.15% for high
889
concentrations. This extremely low difference has no practical impact on the final
890
results and the utilization of equation 4 is acceptable. Moreover, equation 40 is valid
891
up to 60oC, something that maybe makes it unsuitable for the present study. In any
892
case, it is proved that a deviation in viscosity of about 20% (maximum deviation)
893
leads to an extreme low deviation in the PTC thermal efficiency (about 0.15%).
AC C
EP
TE D
M AN U
SC
RI PT
874
50
ACCEPTED MANUSCRIPT Nanofluid - T = 300˚C (equation 4a) 0.00080
Nanofluid - T = 300˚C (equation 40)
0.00075 0.00065 0.00055 0.00050 0.00045 0.00040 0.00
0.01
0.02
0.03
φ
0.05
0.06
Figure 23. Nanofluid dynamic viscosity with two different models
896
AC C
EP
TE D
895
0.04
M AN U
894
RI PT
0.00060
SC
μ (Pa s)
0.00070
897
Figure 24. Comparison of collector thermal efficiency for different viscosity models
898
with inlet temperature equal to 300oC
899 900 51
ACCEPTED MANUSCRIPT 4. Conclusions
902
This work examines the impact of nanofluids in the performance of a solar driven
903
Organic Rankine Cycle with parabolic trough collectors. Four different nanoparticles
904
(Al2O3, CuO, TiO2, Cu) are tested in the thermal oil (Syltherm 800) which is used in
905
the solar field loop. A storage tank with thermal oil is used for storing the heat and
906
this thermal fluid carries the heat to the heat recovery system of the ORC. In the ORC,
907
a regenerative system is used and four different organic working fluids are tested
908
(toluene, MDM, cyclohexane, n-pentane). The system is optimized by selecting the
909
proper nanoparticle concentration and the ideal pressure at the turbine inlet (with the
910
pressure ratio parameter) for all the combinations of nanofluids and organic fluids.
911
The main conclusions of this study are the following:
912
- Toluene is the best organic working fluid, with MDM, cyclohexane and n-pentane to
913
follow respectively. According to the optimization results, CuO is the best candidate,
914
while Cu, Al2O3 and TiO2 are following respectively. Moreover, it is proved that all
915
the nanofluids perform better than the pure thermal oil case.
916
- In the preliminary analysis for the solar field without the ORC, Cu is proved to be
917
the best nanoparticle. This difference between the preliminary analysis and the
918
optimization of the total system indicates that the optimization should be made for the
919
total system and not for each part separately.
920
- The best case is the one with toluene and CuO nanoparticle inside the Syltherm 800.
921
The optimum concentration is found to be 3.98% and the optimum pressure ratio is
922
75.51%. In this case, the electricity production is 167.05 kW and the system
923
efficiency is found to be 20.11%.
AC C
EP
TE D
M AN U
SC
RI PT
901
52
ACCEPTED MANUSCRIPT - The electricity production enhancement for the optimum case, compared to the
925
optimized case with pure thermal oil as working fluid in the solar field, is found to be
926
1.75%.
927
- It is found that the most efficient organic working fluids (toluene and MDM) need
928
relatively lower concentration in order to achieve maximum system performance.
929
Furthermore, it is proved that Al2O3 and CuO have to be used in higher
930
concentrations, compared to TiO2 and Cu.
931
- The results of this work indicate that the use of nanofluids as working fluids in
932
concentrated collectors (PTC) can enhance the system performance, achieving
933
enhancement up to 1.75% compared to operation with pure thermal oil. This fact
934
encourages the utilization of nanofluids in applications as the solar driven Organic
935
Rankine Cycles in order to achieve higher performances.
936
- It is essential to state that the proper working fluid selection is more important than
937
the suitable nanoparticle selection. This result is based on the small margin on the
938
solar thermal efficiency enhancement due to the high performance of PTCs.
939
Nomenclature
940
Aa
Area of the module, m2
941
Ast
Storage tank outer area, m2
942
cp
Specific heat capacity under constant pressure, kJ/kg K
943
D
Diameter, mm
944
df
Equivalent diameter of a base fluid molecule, m
945
dp
Diameter of the nanoparticle, m
AC C
EP
TE D
M AN U
SC
RI PT
924
53
ACCEPTED MANUSCRIPT Gb
Solar beam radiation, kW/m2
947
h
Heat transfer coefficient, W/m2K
948
hout
Convection coefficient between cover and ambient, W/m2K
949
k
Thermal conductivity, W/mK
950
K
Incident angle modifier, -
951
L
Tube length, mm
952
m
Mass flow rate, kg/s
953
MW
Molecular weight, kg/kmol
954
N
Number of modules, -
955
Na
Avogadro number, kmol-1
956
Nu
Mean Nusselt number
957
p
Pressure, bar
958
Pel
Electricity production in the generator, kW
959
Pnet
Net electricity production, kW
960
PP
Pinch point, oC
961
Pr
Prandtl number, -
962
PR
Pressure ratio, -
963
Q
Heat flux, W
964
Re
Reynolds number, -
AC C
EP
TE D
M AN U
SC
RI PT
946
54
ACCEPTED MANUSCRIPT s
Specific entropy, kJ/kgK
966
T
Temperature, oC
967
Ust
Storage tank thermal loss coefficient, kW/m2K
968
(UA)st Heat exchanger effectiveness, kW/m2K
969
V
Volumetric flow rate, m3/h
970
Vst
Volume of storage tank, m3
971
Vwind Ambient wind velocity, m/s
972
Wp
973
Greek symbols
974
β
Nanolayer thickness to the original particle radius, -
975
∆Τrc
Temperature difference in the recuperator, oC
976
ε
Emittance, -
977
ηis
Turbine isentropic efficiency, -
978
ηmg
Electromechanical efficiency of the generator, -
979
ηorc
ORC thermodynamic efficiency, -
980
ηopt
Optical efficiency, -
981
ηpum
Organic fluid pump efficiency, -
982
ηsys
System efficiency, -
983
ηth
Thermal collector efficiency, -
SC M AN U
AC C
EP
TE D
Pump consumption, kW
RI PT
965
55
θ
Incident angle, o
985
µ
Dynamic viscosity, Pa s
986
ρ
Density, kg/m3
987
σ
Stefan–Boltzmann constant [= 5.67 · 10-5 kW/m2 K4]
988
φ
Concentration of nanoparticles, %
989
Subscripts and superscripts
990
am
ambient
991
bf
base fluid
992
c
cover
993
ci
inner cover
994
co
outer cover
995
col
collector
996
col,in collector in
997
col,out collector out
998
con
condenser
999
crit
critical
1000
G
generator
1001
fm
mean fluid
1002
h
high
AC C
EP
TE D
M AN U
SC
984
RI PT
ACCEPTED MANUSCRIPT
56
ACCEPTED MANUSCRIPT in
input
1004
is
isentropic
1005
l
low
1006
loss
losses
1007
max
maximum
1008
nf
nanofluid
1009
np
nanoparticle
1010
o
organic
1011
opt
optimum
1012
r
receiver
1013
ri
inner receiver
1014
ro
outer receiver
1015
s
source – thermal oil
1016
sat
saturation
1017
sky
sky equivalent
1018
sol
solar
1019
st
storage
1020
stored stored quantity
1021
s,in
AC C
EP
TE D
M AN U
SC
RI PT
1003
source in 57
ACCEPTED MANUSCRIPT s,m
source intermediate (in pinch point position)
1023
s,out
source out
1024
sol,t
solar total
1025
u
useful
1026
u,t
useful total
1027
Abbreviations
1028
CFD
Computational Fluid Dynamics
1029
ECO
Economizer
1030
EES
Engineer Equator Solver
1031
EVAP Evaporator
1032
GWP Global Warming Potential
1033
HRS
Heat Recovery System
1034
n.a.
Not available
1035
ODP
Ozone Depletion Potential
1036
ORC Organic Rankine Cycle
1037
PTC
AC C
EP
TE D
M AN U
SC
RI PT
1022
Parabolic Trough Collector
1038
1039
1040 58
ACCEPTED MANUSCRIPT References
1042
[1] Y. Tong, J. Kim, H. Cho, Effects of thermal performance of enclosed-type
1043
evacuated U-tube solar collector with multi-walled carbon nanotube/water nanofluid,
1044
Renewable Energy 2015;83:463-473
1045
[2] D. Zhao, E. Xu, Z. Wang, Q. Yu, L. Xu, L. Zhu, Influences of installation and
1046
tracking errors on the optical performance of a solar parabolic trough collector,
1047
Renewable Energy 2016;94:197-212
1048
[3] T.B. Gorji, A.A. Ranjbar, Thermal and exergy optimization of a nanofluid-based
1049
direct absorption solar collector, Renewable Energy 2017;106:274-287
1050
[4] E. Bellos, C. Tzivanidis, K. Moschos, K.A. Antonopoulos, Energetic and financial
1051
evaluation of solar assisted heat pump space heating systems, Energy Conversion and
1052
Management 2016;120:306-319
1053
[5] A. Zamzamian, M. KeyanpourRad, M. KianiNeyestani, M. Tajik Jamal-Abad, An
1054
experimental study on the effect of Cu-synthesized/EG nanofluid on the efficiency of
1055
flat-plate solar collectors, Renewable Energy 2014;71:658-664
1056
[6] S. Delfani, M. Karami, M.A. Akhavan- Behabadi, Performance characteristics of a
1057
residential-type direct absorption solar collector using MWCNT nanofluid,
1058
Renewable Energy 2016;87(1):754-764
1059
[7] L. Zhang, L. Chen, J. Liu, X. Fang, Z. Zhang, Effect of morphology of carbon
1060
nanomaterials on thermo-physical characteristics, optical properties and photo-
1061
thermal conversion performance of nanofluids, Renewable Energy 2016;99:888-897
AC C
EP
TE D
M AN U
SC
RI PT
1041
59
ACCEPTED MANUSCRIPT [8] S.U. Choi, J. Eastman, Enhancing Thermal Conductivity of Fluids with
1063
Nanoparticles, Argonne National Lab, IL (United States), 1995.
1064
[9] E. Bellos, C. Tzivanidis, Parametric investigation of nanofluids utilization in
1065
parabolic trough collectors, Thermal Science and Engineering Progress 2017;2:71-79
1066
[10] R. Loni, E. Askari Asli-ardeh, B. Ghobadian, A.B. Kasaeian, Sh. Gorjian,
1067
Thermodynamic Analysis of a Solar Dish Receiver using Different Nanofluids,
1068
Energy (2017), doi:10.1016/j.energy.2017.05.016
1069
[11] O. Mahian, A. Kianifar, S.Z. Heris, D. Wen, A.Z. Sahin, S. Wongwises,
1070
Nanofluids effects on the evaporation rate in a solar still equipped with a heat
1071
exchanger, Nano Energy 2017;36:134-155
1072
[12] M. Nourani, N. Hamdami, J. Keramat, A. Moheb, M. Shahedi, Preparation of a
1073
stable nanocomposite phase change material (NCPCM) using sodium stearoyl
1074
lactylate (SSL) as the surfactant and evaluation of its stability using image analysis,
1075
Renewable Energy 2016;93:404-411
1076
[13] J. Liu, F. Wang, L. Zhang, X. Fang, Z. Zhang, Thermodynamic properties and
1077
thermal stability of ionic liquid-based nanofluids containing graphene as advanced
1078
heat transfer fluids for medium-to-high-temperature applications, Renewable Energy
1079
2014;63:519-523
1080
[14] A. Mwesigye, Z. Huan, Thermodynamic analysis and optimization of fully
1081
developed turbulent forced convection in a circular tube with water–Al2O3 nanofluid,
1082
International Journal of Heat and Mass Transfer 2015;l89:694-706
AC C
EP
TE D
M AN U
SC
RI PT
1062
60
ACCEPTED MANUSCRIPT [15] A. Mwesigye, Z. Huan, J.P. Meyer, Thermodynamic optimisation of the
1084
performance of a parabolic trough receiver using synthetic oil–Al2O3 nanofluid,
1085
Applied Energy 2015;156:398-412
1086
[16] A. Mwesigye, Z. Huan, J.P. Meyer, Thermal performance and entropy generation
1087
analysis of a high concentration ratio parabolic trough solar collector with Cu-
1088
Therminol®VP-1 nanofluid, Energy Conversion and Management 2016;120:449-465
1089
[17] T. Sokhansefat, A.B. Kasaeian, F. Kowsary, Heat transfer enhancement in
1090
parabolic trough collector tube using Al2O3/synthetic oil nanofluid, Renewable and
1091
Sustainable Energy Reviews 2014;33:636-644
1092
[18] P.M. Zadeh, T. Sokhansefat, A.B. Kasaeian, F. Kowsary, A. Akbarzadeh, Hybrid
1093
optimization algorithm for thermal analysis in a solar parabolic trough collector based
1094
on nanofluid, Energy 2015;82:857-864
1095
[19] E. Bellos, C. Tzivanidis, K.A. Antonopoulos, G. Gkinis, Thermal enhancement
1096
of solar parabolic trough collectors by using nanofluids and converging-diverging
1097
absorber tube, Renewable Energy 2016;94:213-222
1098
[20] Z. Said, R. Saidur, M.A. Sabiha, A. Hepbasli, N.A. Rahim, Energy and exergy
1099
efficiency of a flat plate solar collector using pH treated Al2O3 nanofluid, Journal of
1100
Cleaner Production 2016;112(5):3915–3926
1101
[21] T. Yousefi, F. Veysi, E. Shojaeizadeh, S. Zinadini, An experimental investigation
1102
on the effect of Al2O3–H2O nanofluid on the efficiency of flat-plate solar collectors,
1103
Renewable Energy 2012;39(1):293-298
AC C
EP
TE D
M AN U
SC
RI PT
1083
61
ACCEPTED MANUSCRIPT [22] S.E. Ghasemi, A.A. Ranjbar, Thermal performance analysis of solar parabolic
1105
trough collector using nanofluid as working fluid: A CFD modelling study, Journal of
1106
Molecular Liquids 2016;222:159-166
1107
[23] A. Kasaeian, S. Daviran, R.D. Azarian, A. Rashidi, Performance evaluation and
1108
nanofluid using capability study of a solar parabolic trough collector, Energy
1109
Conversion and Management 2015;89:368-375
1110
[24] A. Risi, M. Milanese, D. Laforgia, Modelling and optimization of transparent
1111
parabolic trough collector based on gas-phase nanofluids, Renewable Energy
1112
2013;58:134-139
1113
[25] A.E. Kabeel, E.M.S. El-Said, A hybrid solar desalination system of air
1114
humidification,
1115
Experimental investigation, Desalination 2014;341:50-60
1116
[26] N.H. Abu-Hamdeh, K.H. Almitani, Solar liquid desiccant regeneration and
1117
nanofluids in evaporative cooling for greenhouse food production in Saudi Arabia,
1118
Solar Energy 2016;134:202-210
1119
[27] F.A. Boyaghchi, M. Mahmoodnezhad, V. Sabeti, Exergoeconomic analysis and
1120
optimization of a solar driven dual-evaporator vapor compression-absorption cascade
1121
refrigeration system using water/CuO nanofluid, Journal of Cleaner Production
1122
2016;139:970-985, Volume 139
1123
[28] F.A. Boyaghchi, M. Chavoshi, V. Sabeti, Optimization of a novel combined
1124
cooling, heating and power cycle driven by geothermal and solar energies using the
1125
water/CuO (copper oxide) nanofluid, Energy 2015;91:685-699
and
water
flashing
evaporation:
Part
II.
AC C
EP
TE D
dehumidification
M AN U
SC
RI PT
1104
62
ACCEPTED MANUSCRIPT [29] F.A. Boyaghchi, M. Chavoshi, Multi-criteria optimization of a micro solar-
1127
geothermal CCHP system applying water/CuO nanofluid based on exergy,
1128
exergoeconomic and exergoenvironmental concepts, Applied Thermal Engineering
1129
2017;112:660-675
1130
[30] A. Alashkar, M. Gadalla, Thermo-economic analysis of an integrated solar power
1131
generation system using nanofluids, Applied Energy 2017;191:469-491
1132
[31] S. Toghyani, E. Baniasadi, E. Afshari, Thermodynamic analysis and optimization
1133
of an integrated Rankine power cycle and nano-fluid based parabolic trough solar
1134
collector, Energy Conversion and Management 2016;121:93-104
1135
[32] A. Al-Sharafi, A.Z. Sahin, B.S. Yilbas, Overall performance index for hybrid
1136
power plants, Energy Conversion and Management 2015;100:103-106
1137
[33] C. Tzivanidis, E. Bellos, K.A. Antonopoulos, Energetic and financial
1138
investigation of a stand-alone solar-thermal Organic Rankine Cycle power plant,
1139
Energy Conversion and Management 2016;126:421-433
1140
[34] EuroTrough: Development of a Low Cost European Parabolic Trough Collector -
1141
EuroTrough. Final Report, Research funded in part by The European Commission in
1142
the framework of the Non-Nuclear Energy Programme JOULE III. Contract JOR3-
1143
CT98-0231;2001.
1144
[35] http://www.loikitsdistribution.com/files/syltherm-800-technical-data-sheet.pdf
1145
[36] M. Turkyilmazoglu, Condensation of laminar film over curved vertical walls
1146
using single and two-phase nanofluid models, European Journal of Mechanics -
1147
B/Fluids 2017;65:194-191
AC C
EP
TE D
M AN U
SC
RI PT
1126
63
ACCEPTED MANUSCRIPT [37] F-Chart Software, Engineering Equation Solver (EES); 2015. Available at:
1149
http://www. fchart.com/ees.
1150
[38] M. Ayatollahi, S.H. Nasiri, A.B. Kasaeian, Convection heat transfer modeling of
1151
Ag nanofluid using different viscosity theories, IIUM Engineering Journal
1152
2012;13(1):1-11
1153
[39] K. Khanafer, K. Vafai, A critical synthesis of thermophysical characteristics of
1154
nanofluids, International Journal of Heat and Mass Transfer 2011;54:4410-4428
1155
[40] W. Yu, S.U.S. Choi, The role of interfacial layers in the enhanced thermal
1156
conductivity of nanofluids: a renovated Maxwell model, Journal of Nanoparticle
1157
Research 2003;5:167
1158
[41] W. Duangthongsuk, S. Wongwises, An experimental study on the heat transfer
1159
performance and pressure drop of TiO2-water nanofluids flowing under a turbulent
1160
flow regime, International Journal of Heat and Mass Transfer 2010;53(1-3):334-344
1161
[42] G.K. Batchelor, The effect of Brownian motion on the bulk stress in a suspension
1162
of spherical particles, J. Fluid Mech. 1977;83:97-117
1163
[43] O. Mahian, S. Mahmud, S.Z. Heris, Effect of Uncertainties in Physical Properties
1164
on Entropy Generation Between Two Rotating Cylinders With Nanofluids, Journal of
1165
Heat Tranfer 2012;134: 101704-1
1166
[44] O. Mahian, A. Kianifar, A.Z. Sahin, S. Wongwises, Entropy generation during
1167
Al2O3/water nanofluid flow in a solar collector: Effects of tube roughness,
1168
nanoparticle size, and different thermophysical models, International Journal of Heat
1169
and Mass Transfer 2014;78:64-75
AC C
EP
TE D
M AN U
SC
RI PT
1148
64
ACCEPTED MANUSCRIPT [45] C.A.R. Sotomonte, S. Ribeiro, E. Oliveira, E.E.S. Lora, O.J. Venturini, Organic
1171
Rankine cycle associated with an absorption chiller for biomass applications,
1172
Engenharia Térmica (Thermal Engineering) 2011;10(01-02):15-22
1173
[46] J.A. Duffie, W.A. Beckman, Solar Engineering of Thermal Processes, 3rd
1174
edition, Wiley, Hoboken, NJ, USA, 2006
1175
[47] E. Bellos, C. Tzivanidis, K.A. Antonopoulos, I. Daniil, The use of gas working
1176
fluids in parabolic trough collectors – An energetic and exergetic analysis, Applied
1177
Thermal Engineering 2016;109(A);1-14
1178
[48] R. Forristall, Heat transfer analysis and modeling of a parabolic trough solar
1179
receiver implemented in engineering equation solver. Colorado: National Renewable
1180
Energy Laboratory (NREL); 2003.
1181
[49] Y. Qiu, M.-J. Li, Y.-L. He, W.-Q. Tao, Thermal performance analysis of a
1182
parabolic trough solar collector using supercritical CO2 as heat transfer fluid under
1183
non-uniform solar flux, Applied Thermal Engineering 2017;115:1255-1265
1184
[50] W.C. Swinbank, Long-wave radiation from clear skies. QJR Meteorol Soc
1185
1963;89:339–40
1186
[51] E. Bellos, E. Mathioulakis, C. Tzivanidis, V. Belessiotis, K.A. Antonopoulos,
1187
Experimental and numerical investigation of a linear Fresnel solar collector with flat
1188
plate receiver, Energy Conversion and Management 2016;130:44-59
1189
[52] J. Leinhard IV, J. Leinhard V, A Heat Tranfer Textbook, 4th edition. Philogiston
1190
Press, USA, 2012, 354–360
AC C
EP
TE D
M AN U
SC
RI PT
1170
65
ACCEPTED MANUSCRIPT [53] B.C. Pak, Y.I. Cho, Hydrodynamic and heat transfer study of dispersed fluids
1192
with submicron metallic oxide particles, Experimental Heat Transfer 1998;11:151–70.
1193
[54] Y. Xuan, Q. Li, Investigation on Convective Heat Transfer and Flow Features of
1194
Nanofluids. ASME. J. Heat Transfer. 2003;125(1):151-155
1195
[55] E. Bellos, C. Tzivanidis, K.A. Antonopoulos, Exergetic and energetic
1196
comparison of LiCl-H2O and LiBr-H2O working pairs in a solar absorption cooling
1197
system, Energy Conversion and Management 2016;123:453-461
1198
[56] O. Behar, A. Khellaf, K. Mohammedi, A novel parabolic trough solar collector
1199
model – Validation with experimental data and comparison to Engineering Equation
1200
Solver (EES), Energy Conversion and Management 2015;106:268-281
1201
[57] E. Bellos, C. Tzivanidis, Kimon A. Antonopoulos, A detailed working fluid
1202
investigation for solar parabolic trough collectors, Applied Thermal Engineering
1203
2017;114:374-386
1204
[58] M. Corcione, Empirical correlating equations for predicting the effective thermal
1205
conductivity and dynamic viscosity of nanofluids, Energy Conversion and
1206
Management 2011;52:789-793
AC C
EP
TE D
M AN U
SC
RI PT
1191
66