Vous êtes sur la page 1sur 8

Applied Thermal Engineering 80 (2015) 339e346

Contents lists available at ScienceDirect

Applied Thermal Engineering


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

Research paper

Design, construction, and preliminary results of a 250-kW organic


Rankine cycle system
Ben-Ran Fu*, Yuh-Ren Lee, Jui-Ching Hsieh
Green Energy and Environment Research Laboratories, Industrial Technology Research Institute, Hsinchu 31040, Taiwan

h i g h l i g h t s
 A 250-kW ORC system using turbine expander was studied for waste heat recovery.
 The experimentally maximal net power output was 219.5 5.5 kW.
 The experimentally maximal system thermal efciency was 7.94%.
 The turbine isentropic efciency was 63.7% with a rotational speed of 12,386 rpm.
 The system responded very rapidly as the heat source temperature changed.

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 3 December 2014
Accepted 30 January 2015
Available online 11 February 2015

This study involved designing and constructing a 250-kW organic Rankine cycle system, consisting of a
pump, preheater, evaporator, turbine, generator, condenser, as well as hot and cooling water circulation
systems. Refrigerant R245fa was used as a working uid. The design operating pressure levels of the
preheater/evaporator and condenser was 1.265 MPa and 0.242 MPa, respectively. Under design conditions, the net power output was 243 kW and the system thermal efciency was 9.5%. The preliminary
experimental results under off-design conditions showed that the average net power output was
219.5 kW with a uctuation of 5.5 kW during prolonged operation. The maximal net power output and
system thermal efciency were 225 kW and 7.94%, respectively. Under this condition, the isentropic
efciency of the turbine was 63.7% with a rotational speed of 12 386 rpm, and the back-work ratio was
6.7%. In addition, the results of the dynamic testing demonstrated that the present system responded
very rapidly as the heat source temperature changed. The experimental results also demonstrated that
the system thermal efciency and net power output increased linearly with an increasing heat source
temperature. However, the effect of the heat source temperature on the turbine efciency was not
obvious.
2015 Elsevier Ltd. All rights reserved.

Keywords:
Organic Rankine cycle (ORC)
Turbine expander
Thermal efciency
Waste heat recovery

1. Introduction
An organic Rankine cycle (ORC) is identical to a steam Rankine
cycle, except that it employs organic uids with a low boiling point
as working uids to generate power from low-temperature heat
sources [1]. ORC is considered to be one of the most economical and
efcient methods for converting low-grade thermal energy, such as
that derived from waste heat recovery, geothermal and solar thermal sources, biomass combined heat and power (CHP), and ocean
thermal energy into electricity [2,3]. Previous studies on ORCs have

* Corresponding author.
E-mail address: brfu@mx.nthu.edu.tw (B.-R. Fu).
http://dx.doi.org/10.1016/j.applthermaleng.2015.01.077
1359-4311/ 2015 Elsevier Ltd. All rights reserved.

applied various perspectives and research tools, including conducting technical-economic-market surveys [1,4], developing
methods for selecting working uids [5], reviewing application of
scroll expanders for ORC systems [6], evaluating waste heat recovery from a power plant [7], onboard ships [8], and at data
centers [9], as well as proposing proof-of-concepts [10], optimal
control strategy models [11], quasi-dynamic models [12]. In addition, relevant studies have assessed the effect of the optimal pinchpoint temperature range of evaporators on system performance
[13], conducted prototype testing [14e16], and performed statistical analysis of ORC-related patent data [17] and off-design performance analysis [18,19]. This section reviews several previous
experimental studies on the ORC systems in detail.

340

B.-R. Fu et al. / Applied Thermal Engineering 80 (2015) 339e346

Nomenclature
Esys
Etur
mW
s
T
TC,in
TC,out
TW,in
TW,out
Wnet

thermal efciency of the ORC system (%)


isentropic efciency of the turbine (%)
mass ow rate of the hot water (kg/s)
mass specic entropy (kJ/kg K)
temperature ( C)
inlet temperature of the cooling water ( C)
outlet temperature of the cooling water ( C)
inlet temperature of the hot water ( C)
outlet temperature of the hot water ( C)
net power output of the system (kW)

Manolakos et al. [20] experimentally evaluated the performance


of a low-temperature solar ORC system for reverse osmosis (RO)
desalination using R134a as the working uid and scroll expander.
In their experiments, the expander had a maximal efciency of 65%,
power output of 2.05 kW, and ORC efciency of 4%. In addition,
Manolakos et al. [21] presented on-site experimental results of a
solar ORC system combined with an RO system. However, the
maximal efciencies of the expander and the ORC system were
approximately 45% and 1.75%, respectively.
Lemort et al. [22] and Quoilin et al. [23] have experimentally
investigated an ORC prototype with an open-drive oil-free scroll
expander and R123 as the working uid. In their experiments, the
maximal isentropic effectiveness of the expander was 68%, and the
shaft work of the expander was between 0.4 and 1.82 kW. The
maximal system efciency was 7.4%. In addition, they determined
that the deviation between the experimental and predicted results
of the proposed model was only approximately 5%.
Wang et al. [24] reported on the experimental results of an onsite micro-scale solar ORC system using R245fa and R245fa/R152
mixtures as the working uids. The maximal power output and ORC
efciency were 7.2 W and 5.59%, respectively. Their results
demonstrated that the ORC efciency was considerably improved
by using R245fa/R152 mixtures rather than using pure R245fa.
Wang et al. [25] designed and constructed a low-temperature
solar ORC system using R245fa as the working uid and a rollingpiston expander. In their experiments, the average shaft power
output and isentropic efciency of the expander were 1.73 kW and
45.2%, respectively. They achieved a maximal ORC efciency of
12.9%.
Pei et al. [26] experimentally examined a kW-scale ORC system
using a turbine and R123 as the working uid. The results showed
that the isentropic efciency of the turbine was 62.5%; the ORC
efciency was 6.8% at a temperature difference of 70  C between
the heat source and heat sink, and the power output was 1.36 kW.
Qui et al. [27] constructed and tested a micro-scale biomassred CHP system with an ORC. Their experimental results showed
that the micro-CHP system with an ORC generated 0.861 kW and
47.26 kW in electricity and heat, respectively, and the corresponding efciencies of the expander and ORC were 53.92% and
3.78%, respectively. Consequently, the overall biomass CHP efciency was 78.69%.
Zheng et al. [28] proposed a kW-scale rolling-piston expander
for a low-temperature ORC system using R245fa as the working
uid and conducted a running test of the proposed ORC system. The
experimental results showed that the expander operated at
350e800 rpm with a maximal power output of 0.35 kW when the
heat source temperature was below 90  C. The maximal isentropic
efciency of the expander and cycle efciency were 43.3% and 5%,
respectively.

Declaye et al. [29] experimentally investigated an ORC system


using R245fa as the working uid and a scroll expander. They
demonstrated a maximal isentropic efciency of the expander and
shaft power of 75.7% and 2.1 kW, respectively. A maximal cycle
efciency of 8.5% was reached at evaporating and condensing
temperatures of 97.5 and 26.6  C, respectively.
Twomey et al. [30] reported on the dynamic performance of a
small-scale solar cogeneration system with an ORC using R134a as
the working uid and a scroll expander. The results demonstrated a
maximal isentropic efciency of the expander of 59% with a corresponding ORC efciency of 3.47%. In addition, the maximal power
output was 0.676 kW.
Hsu et al. [16] experimentally investigated the effect of inlet
pressure and the pressure ratio in the expander on the performance
of a 50-kW ORC system using a screw expander and R245fa as the
working uid. The results showed that for a given pressure ratio,
higher inlet pressure resulted in both higher isentropic efciency of
the expander and higher system efciency under an overexpansion condition, but resulted in lower isentropic efciency of
the expander and system efciency under an under-expansion
condition. In their experiments, the maximal power output was
50 kW and the system thermal efciency was 10.5%.
Jradi and Riffat [31] experimentally examined a small-scale trigeneration system, consisting of an ORC-based (using a scroll
expander and HFE7100 as the working uid) CHP unit and a combined dehumidication and cooling unit. They demonstrated that
this combined system provided approximately 9.6, 6.5, and 0.5 kW
for heating, cooling, and electric power, respectively. Under a
maximal electric power output condition, the isentropic efciency
of the expander was 74.2% and the corresponding cycle efciency
was 5.64%.
Avadhanula and Lin [32] proposed empirical models for a screw
expander based on the experimental data of a 50-kW ORC system
using R245fa as the working uid. The experimental results showed
that the pressure ratio, volume ratio, and power output of the ORC
system were 2.70e6.54, 2.54e6.20, and 10e51.5 kW, respectively.
In addition, their proposed models predicted the system power
output accurately, namely within 10% and 7.5% of experimental
values for the polytropic exponent model and isentropic work
output model, respectively. The maximal isentropic efciency of
the expander was nearly 70%, but no information of the ORC efciency was provided.
Klonowicz et al. [33] designed and constructed a small-scale
turbine with a rotational speed of 3264 rpm for an ORC system using R227ea as the working uid. In addition, they presented the
numerical and preliminary testing results of the system. The
experimental power output was 9.9 0.2 kW, and the corresponding electrical efciency was 53 2%. However, the ORC efciency was not provided. In addition, the predicted system efciency
exhibited a high consistency with the experimental result.
Chang et al. [34] experimentally investigated an ORC system
with a scroll expander and R245fa as the working uid. The
maximal shaft power output and electricity power output were 1.74
and 1.375 kW, respectively, at a temperature difference of 60.6  C
between the heat source and heat sink. The maximal system efciency of the ORC system was 7.77%.
Zhang et al. [35] reported on the experimental results of an ORC
system with a single-screw expander and R123 as the working uid
for waste heat recovery from the exhaust of a diesel engine with a
maximal horsepower of 336 kW. The results demonstrated that the
maximal ORC power output was 10.38 kW. In addition, the ORC
efciency and overall system efciency were 6.48% and 43.8%,
respectively, resulting in 250 kW of diesel engine output. Using the
ORC improved the overall efciency of the diesel engine by
approximately 1.53%.

B.-R. Fu et al. / Applied Thermal Engineering 80 (2015) 339e346

More detailed experimental results of ORC systems from previous studies are summarized in Table 1. The cited studies experimentally investigated ORC systems with a power output of less
than 50 kW. In addition, although numerous refrigerants were
examined in numerous studies (e.g., on the selection of the working
uid in an ORC system [5,36]), the typically used working uids in
the experimental ORC systems were R123, R134a, and R245fa.
Table 1 indicates that scroll expanders were generally used in lowkW-level ORC systems, which is due to the compact size, low cost,
high efciency, and low number of moving parts of scroll expanders
[30]; screw-type expanders were used in tens-of-kW-level ORC
systems; the turbine expander type was generally used in
hundreds-of-kW-level ORC systems. However, based on literature
review, no studies have reported on experimental results of large
ORC systems (>100 kW) using turbine expanders. As shown in
Table 1, ORC thermal efciency was generally lower than 10%.
The present study constructed a 250-kW ORC system (using
R245fa as the working uid), consisting of a pump, preheater,
evaporator, turbine, generator, condenser, as well as hot and cooling water circulation systems. In addition, this paper presents the
preliminary results of the present ORC system under off-design

341

Fig. 1. Photograph of the 250-kW ORC prototype (pump is located at back side). (1)
Preheater, (2) evaporator, (3) turbine, (4) generator, (5) condenser, (6) pump, (A)
cooling water inlet, (B) cooling water outlet, (C) hot water inlet, (D) hot water outlet.

conditions. The effect of the heat source temperature on system


performance was investigated.
2. System design and construction

Table 1
Some experimental results of the ORC system from available literature (sorted by
expander type).
Authors

Expander type
(max. efciency, %)

Wang et al. [24]

Working
uid

R245fa/
R152a
Qiu et al. [27]
Multi-vane (53.92) HFE7000
Peris et al. [40]
Volumetric (65.33) R245fa
Wang et al. [25]
Rolling-piston (45.2) R245fa
Zheng et al. [28]
Rolling-piston (43.3) R245fa
Manolakos et al. [20] Scroll (65)
R134a
Manolakos et al. [21] Scroll (45)
R134a
Lemort et al. [22]
Scroll (68)
R123
Declaye et al. [29]
Scroll (75.7)
R245fa
Twomey et al. [30]
Scroll (59)
R134a
Jradi and Riffat [31] Scroll (74.2)
HFE7100
Chang et al. [34]
Scroll (76)
R245fa
Peterson et al. [41]
Scroll (49.9)
R123
Wang et al. [42]
Scroll (77.5)
R123
Mathias et al. [43]
Scroll (83)
R123
Lemort et al. [44]
Scroll (71.03)
R245fa
Bracco et al. [45]
Scroll (74)
R245fa
Liu et al. [46]
Scroll (41)
R123
Zhou et al. [47]
Scroll (57)
R123
Saitoh et al. [48]
Scroll (65)
R113
Tarique et al. [49]
Scroll (64)
R134a
Li et al. [50]
Scroll (83)
R245fa/
R601a
Miao et al. [51]
Scroll (81)
R123
Gao et al. [52]
Scroll (55.3)
R245fa
Lee et al. [14,15]
Screw (65)
R245fa
Hsu et al. [16]
Screw (72.5)
R245fa
Avadhanula
Screw (70)
R245fa
and Lin [33]
Zhang et al. [35]
Screw (57.88)
R123
Pei et al. [26]
Turbine (62.5)
R123
Klonowicz et al. [33] Turbine (59)
R227ea
Nguyen et al. [53]
Turbine (49.8)
n-Pentane
Liu et al. [54]
Turbine ( )
HFE7000,
HFE7100
Li et al. [55]
Turbine (53)
R123
Kang [56]
Turbine (82.2)
R245fa
Li et al. [57]
Turbine (68)
R123
Borsukiewicz-Gozdur Turbine ( )
R227ea
[58]

Max.
electricity
or shaft
power (kW)
0.0072

Max. ORC
thermal
efciency
(%)
5.59

0.861
15.93
1.73
0.35
2.05
1
1.82
2.1
0.676
0.5
1.375
0.256
0.625
2.96
2.032
1.5
0.76
0.645
0.35
0.92
0.55

3.78
10.88
12.9
5
4
1.75
7.4
8.5
3.47
5.64
7.77
7.2
e
e
e
8
2.9
8.5
11
8.5
4.45

3.25
0.151
50
50
51.5

6.39
3.2
8.05
10.5
e

10.38
1.36
9.9
1.5
0.284

6.48
6.8
e
4.3
e

6.57
32.7
6
9.87

e
5.22
7.98
4.88

This 250-kW ORC prototype, as shown in Fig. 1, was built at the


Industrial Technology Research Institute, Taiwan. The ORC unit
measured 450 cm (length)  270 cm (width)  310 cm (height),
weighing approximately 11,000 kg. The working uid was refrigerant R245fa, one of the most appropriate uids for low-grade
waste heat recovery of ORC systems [37]; the mass ow rate was
11.58 kg/s. The thermodynamic properties of R245fa were evaluated using REFPROP 9.0 [38], which was developed by the National
Institute of Standards and Technology (USA). The working uid
R245fa circulated on the shell side of the heat exchangers, and the
hot and cooling water circulated in the tube side. Fig. 2 depicts the
detailed scheme of the experimental test rig of the system.
2.1. Design conditions
Fig. 3 shows the Tes diagram of the present ORC system. The
design operating pressure levels in the preheater/evaporator and
condenser were 1.265 MPa (the evaporation/saturation temperature was 100  C) and 0.242 MPa (the condensation temperature
was 39  C), respectively. The design points of the heat source (hot
water) temperature and mass ow rate were 133.9  C and 15.39 kg/
s, respectively. The details of mathematical model for each
component and system performance were described in the previous studies [18,19], which were preliminary analyses of the offdesign performance of the ORC system investigated in this study.
Under design conditions, the net power output and system thermal
efciency were 243 kW and 9.5%, respectively.

Fig. 2. Detailed scheme of the experimental test rig.

342

B.-R. Fu et al. / Applied Thermal Engineering 80 (2015) 339e346

Temperature ( C)

hot water

Table 2
Detailed parameters of the used heat exchangers.

(a) Preheater

3
2

2s
1

5s
cooling water

Mass specific entropy (kJ/kgK)


Fig. 3. Tes diagram of the studied ORC system.

2.2. Heat exchangers


The present ORC system featured three heat exchangers, namely
a preheater, an evaporator, and a condenser. The preheater,
comprising 200 heat transfer tubes, was of a one-pass shell-andtube type, in which the heat transfer tube was of a ried type in the
inside and low-nned type in the tube outside. The evaporator,
consisting of 300 heat transfer tubes, was of a four-pass shell-andtube ooded type. The condenser, composed of 480 heat transfer
tubes, was of a two-pass shell-and-tube ooded type. The detailed
parameters of the used preheater, evaporator, and condenser are
summarized in Table 2.
2.3. Pump, turbine, and generator
The pump used in the present system was manufactured by
Grundfos (model: CR 32-5, 60 Hz). The 250-kW turbine was
designed at the Industrial Technology Research Institute (Taiwan)
and by a SoftInWay Inc. (USA) engineering team, and fabricated at
the Industrial Technology Research Institute. Detailed information
on the turbine performance was provided in a previous study [39].
The manufactured blade row is shown in Fig. 4a, and the rotor
wheel mounted with a stator ring and shaft is shown in Fig. 4b. The
rotational speed of the turbine was 12,000 10% rpm with an
isentropic efciency of approximately 80%. The shaft power
generated by the turbine was transferred to the generator through a
gearbox at a rotational speed of 3600 rpm; the generator efciency
was 90%. To prevent leakage of the working uid, the turbine,
gearbox, and generator were arranged in a hermetic space. This
type of hermetic turbogenerator was also employed by Klonowicz
et al. [33].
2.4. Hot and cooling water circulation systems
Hot water, considered the waste heat source, was supplied by
the pressurized hot water boiler with a maximal capacity of
3788 kW, as shown in Fig. 5a. In the currently proposed system,
natural gas was used as a fuel for the boiler. Heated water was
collected in the container, and owed into the ORC system through
the piping. The cooling water circulation system consisted of two
500-RT cooling towers and had a total maximal capacity of
3860 kW, as shown in Fig. 5b.
2.5. Measurement and control
Pressure transducers with the full scale range of 2500 kPa,
manufactured by WIKA (model A-10, Germany), and Pt100 (three
wires) thermometers, designed by MorShine (Taiwan), were
located at each point of the system (i.e., Points 1 to 5 in Fig. 3). The

Tube inside/outside diameter


Tube thickness
Tube number
Tube in window
Tube bundle
Tube inside type
Tube outside type/Fin per inch (FPI)
Tube arrangement
Tube pitch transverse
Tube pitch longitudinal
Tube/Shell length
Shell inside diameter
Bundle hole diameter
Bundle diameter
Sealing strips number
Nozzle inside diameter
Bafe plate diameter
Bafe thickness
Bafe spacing
Bafe cut
Bafe plate number

1.471/1.587 cm
0.058 cm
200
83
1 pass
Ried
Low-nned/42
Staggered
1.984 cm
1.718 cm
360 cm
32.45 cm
1.61 cm
31.66 cm
0
10 cm
31.95 cm
0.4 cm
20 cm
30%
17

(b) Evaporator
Tube inside/outside diameter
Tube thickness
Tube number
Tube bundle
Tube inside type
Tube outside type
Fin height
Fin thickness
Tube arrangement
Tube pitch transverse
Tube pitch longitudinal
Tube/Shell length
Distance upper row/center
Tube in upper row
Number of tube rows
Free space above upper tube row
Shell inside diameter

1.639/1.765 cm
0.063 cm
300
4 pass
Ried
Low-nned
0.06 cm
0.02 cm
Staggered
2.375 cm
2.057 cm
360 cm
1.9 cm
28
12
47%
69.59 cm

(c) Condenser
Tube outside diameter
Tube number
Tube bundle
Tube inside type
Tube outside type
Tube arrangement
Tube pitch transverse
Tube pitch longitudinal
Tube/Shell length
Distance upper row/center
Tube in upper row
Number of tube rows
Shell inside diameter

1.905 cm
480
2 pass
Ried
Low-nned
Staggered
2.375 cm
2.090 cm
360 cm
2.23 cm
23
18
71.7 cm

mass ow rate of the working uid was measured at Point 2 using a


vortex ow meter made by BNC (model 1100/LRT-B/A, Taiwan). In
addition, level meters were employed in the evaporator and
condenser. The rotational speed of the turbine was measured using
a tachometer sensor. The mass ow rate and temperature of the hot
and cooling water were measured using an electromagnetic ow
meter made by BNC (model BMS1000-A100, Taiwan) in the hot and
cooling water circulation systems. The electric power of the induction generator was measured using a power analyzer made by
HIOKI (model 3169-20 with a 9661 clamp on sensor, Japan). The
overall system control (including the processes of startup, shutdown, and emergency turn off) and data acquisition were performed using an AB MicroLogix 1500 programmable logic
controller (USA). The measured data were then transmitted to a

B.-R. Fu et al. / Applied Thermal Engineering 80 (2015) 339e346

343

Fig. 4. (a) Manufactured rotor wheel and (b) rotor wheel mounted with stator ring and shaft.

personal computer for detailed system analysis. The measurement


uncertainties in temperature, pressure, mass ow rate of the
working uid, mass ow rate of the water, heat transfer rate, turbine efciency, system thermal efciency, and net power output are
0.25  C, 1.0%, 0.7%, 0.5%, 1.72%, 1.48%, 3.10%, and 2.36%,
respectively.
3. Preliminary results and discussion
Fig. 6 shows the measured inlet and outlet temperatures of the
hot and cooling water and the system net power output (Wnet)
during prolonged operation. In this gure, TW,in and TW,out are the
inlet and outlet temperatures of the hot water, respectively; and
TC,in and TC,out are the inlet and outlet temperatures of the cooling
water, respectively. The ORC system required approximately 30 min
(excluding the period for the boiler to warm up before the experiments) to reach a steady-state condition and was longer than that
(approximately 15 min) required for the 50-kW ORC system in our
previous study [14]. When the system was under a steady-state
condition, the average inlet temperature of hot water was
119.2  C; however, this was lower than it should have been according to the design (i.e., 133.9  C). This was due to our infrastructure limiting the supply of natural gas. Therefore, during the

experiments, we increased the mass ow rate of the hot water from


15.39 kg/s to 26.70 kg/s to enable the heat transfer rate to meet the
requirement. In addition, because of a lower heat source temperature, a higher evaporation temperature (106  C) was applied to
maintain the power output level as design value.
Fig. 6 also shows that the uctuations of the heat source and
sink temperatures were considerably low (1  C), indicating that
the boiler and cooling towers were functioning adequately; the
average net power output was 219.5 kW with a uctuation of
5.5 kW (approximately 2.5% of the average net power output). The
maximal net power output and system thermal efciency were
225 kW and 7.94%, respectively. Under the maximal net power
output condition, the isentropic efciency of the turbine was 63.7%
with a rotational speed of 12,386 rpm. The pump and turbine efciencies were lower than the designed value; the generator efciency reached 91.5%. In addition, the back-work ratio, dened as
the ratio between the works consumed by the pump and produced
by the expander, was 6.7%. The heat transfer rates in the preheater
and evaporator were 1161 and 1560 kW, respectively, close to the
designed values. Heat loss in the preheater and evaporator was
113 kW, which was approximately 4% of the heat input from the hot
water. The temperatures of two particular cycle points, namely the
superheat at the evaporator outlet and the subcooling at the

Fig. 5. Hot and cooling water circulation systems. (a) Hot water system, (b) two 500-RT cooling towers.

B.-R. Fu et al. / Applied Thermal Engineering 80 (2015) 339e346

250

250

Wnet

150

TW,in

100

100

TW,out

TC,out

50
0

60

120

180

240

300

TW,in
200

Wnet

100

0
360

80

150

TW,out

90

50

TC,in

250

110

T ( C)

T ( C)

150

200

Wnet (kW)

200

120

(a)

(b)

60

(c)

(d)

120

Wnet (kW)

344

100
240

180

Time (min)

Time (min)

Fig. 6. Temperatures of hot and cooling water and system net power output.

Fig. 7. Dynamic behavior of the system.

condenser outlet, which were not considered in the previous


analysis of the design condition, were 1.7 and 2.5  C, respectively. A
slight degree of superheat was applied to ensure that no liquidphase working uid entered the turbine. The detailed experimental results at the maximal power output and design parameters
are summarized in Table 3.
After prolonged operation, a dynamic testing was also conducted. Fig. 7 shows the dynamic behavior of the system under the
continued change in the inlet temperature of the heat source, i.e.,
TW,in. During the dynamic testing, there were 4 periods, which the
operation conditions were described as follows: (a) TW,in decreased
rapidly from the steady-averaged value of 119.2  Ce102.7  C; (b)
TW,in increased rapidly from 102.7  C to 115.0  C; (c) TW,in decreased
slowly from 115.0  C to 102.7  C; and (d) TW,in increased slowly from
102.7  C to 109.4  C. The corresponding dynamic responses of the
system net power output (Wnet) were described as follows: (a) Wnet
decreased rapidly from 223.4 kW to 132.0 kW; (b) Wnet increased
rapidly from 132.0 kW to 197.9 kW; (c) Wnet decreased slowly from
197.9 kW to 135.8 kW; and (d) Wnet increased slowly from 135.8 kW
to 169.7 kW. In addition, the outlet temperature of the heat source,
i.e., TW,out, shows the same dynamic behavior with Wnet. Most
importantly, those dynamic results demonstrated that the present
system responded very rapidly as the heat source temperature
changed.

Furthermore, we studied the effect of the heat source temperature (TW,in) on system performance, as shown in Fig. 8. The
experimental results show that as TW,in increased from 102.4 to
119.8  C, the isentropic efciency of the turbine slightly increased
from 58.8% to 63.7%; the net power output increased linearly and
substantially from 135 to 225 kW. We observed such a linear increase of net power output in our previous off-design analysis [18].
Furthermore, the system thermal efciency increased from 6.31% to
7.94% for the studied range of TW,in, indicating that the overall
thermal efciency increased by 0.94%, as the heat source temperature increased by 10  C. The increase in thermal efciency was
consistent with previous analytical results [18], which demonstrated a 0.91% efciency increase for a 10  C of TW,in.

Design

Experiments

Mass ow rate of hot water (kg/s)


Inlet temperature of hot water ( C)
Outlet temperature of hot water ( C)
Mass ow rate of R245fa (kg/s)
Evaporation temperature ( C)
Evaporation pressure (MPa)
Superheat at evaporator outlet ( C)
Total heat input from hot water (kW)
Heat transfer rate in the preheater (kW)
Heat transfer rate in the evaporator (kW)
Heat loss in the preheater and evaporator (kW)
Condensation temperature ( C)
Condensation pressure (MPa)
Subcooling at condensation outlet ( C)
Rotational speed of turbine (rpm)
Back work ratio (%)
Pump efciency (%)
Turbine efciency (%)
Generator efciency (%)
ORC thermal efciency (%)
Net power output (kW)

15.39
133.9
94.4
11.58
100
1.265
0
2573
1016
1557
0
39
0.242
0
12,000
4.1
90
80
90
9.5
243

26.70
119.8
94.7
11.85
106
1.440
1.7
2834
1161
1560
113
40
0.251
2.5
12386
6.7
68.1
63.7
91.5
7.94
225

250

Wnet

200

Esys

150
7
100

Etur

50
0
100

105

110

115

120

Esys (%)

Parameter

The present study involved designing and constructing a 250kW ORC system, consisting of a pump, preheater, evaporator, turbine, generator, condenser, as well as hot and cooling water circulation systems. The ORC unit measured 450 cm (length)  270 cm
(width)  310 cm (height), weighing approximately 11,000 kg.
Refrigerant R245fa was used as the working uid. The design
operating pressure levels of the preheater/evaporator and
condenser were 1.265 MPa and 0.242 MPa, respectively. Under
design conditions, the net power output was 243 kW and the system thermal efciency was 9.5%.
The preliminary experimental results of the ORC system under
off-design conditions showed that the average net power output
was 219.5 kW with a uctuation of 5.5 kW during prolonged

Wnet (kW) or Etur (%)

Table 3
Detailed experimental results at the maximum net power output.

4. Conclusions

5
125

TW,in ( C)
Fig. 8. Net power output, turbine and system efciencies as a function of heat source
temperature.

B.-R. Fu et al. / Applied Thermal Engineering 80 (2015) 339e346

operation. The maximal net power output and system thermal efciency were 225 kW and 7.94%, respectively, at evaporation and
condensation temperatures of 106 and 40  C. Under this condition,
the isentropic efciency of the turbine was 63.7% with a rotational
speed of 12,386 rpm, and the back-work ratio was 6.7%. In addition,
the dynamic behavior of the system under the continued change in
the inlet temperature of the heat source was also studied. The results of the dynamic testing demonstrated that the present system
responded very rapidly as the heat source temperature changed.
The experimental results also demonstrated that the system thermal efciency and net power output increased linearly and substantially with an increase of the heat source temperature.
However, the effect of the heat source temperature on the turbine
efciency was not obvious.
Future study will focus on obtaining more experimental results
and constructing a map of system performance under off-design
heat source conditions. Such research could provide operational
guidelines and indicate an optimal control strategy for off-design
operation. In addition, we will also develop an articialintelligence feedback control system for the ORC unit.
Acknowledgements
The authors express their gratitude for the Energy R&D foundation funding from the Bureau of Energy of the Ministry of Economic Affairs, Taiwan, under the grant number of 104-E0207.
References
[1] S. Quoilin, M.V.D. Broek, S. Declaye, P. Dewallef, V. Lemort, Techno-economic
survey of organic Rankine cycle (ORC) systems, Renew. Sustain. Energy Rev.
22 (2013) 168e186.
[2] B.F. Tchanche, Gr Lambrinos, A. Frangoudakis, G. Papadakis, Low-grade heat
conversion into power using organic Rankine cycles e a review of various
applications, Renew. Sustain. Energy Rev. 15 (2011) 3963e3979.
[3] J. Wang, Z. Yan, M. Wang, S. Maa, Y. Dai, Thermodynamic analysis and optimization of an (organic Rankine cycle) ORC using low grade heat source,
Energy 49 (2013) 356e365.
lez, J.J. Segovia, M.C. Martn, G. Antoln, F. Chejne, A. Quijano, A technical,
[4] F. Ve
economical and market review of organic Rankine cycles for the conversion of
low-grade heat for power generation, Renew. Sustain. Energy Rev. 16 (2012)
4175e4189.
[5] J. Bao, L. Zhao, A review of working uid and expander selections for organic
Rankine cycle, Renew. Sustain. Energy Rev. 24 (2013) 325e342.
[6] P. Song, M. Wei, L. Shi, S.N. Danish, C. Ma, A review of scroll expanders for
organic Rankine cycle systems, Appl. Therm. Eng. 75 (2015) 54e64.
[7] D. Gewald, K. Siokos, S. Karellas, H. Spliethoff, Waste heat recovery from a
landll gas-red power plant, Renew. Sustain. Energy Rev. 16 (2012)
1779e1789.
[8] G. Shu, Y. Liang, H. Wei, H. Tian, J. Zhao, L. Liu, A review of waste heat recovery
on two-stroke IC engine aboard ships, Renew. Sustain. Energy Rev. 19 (2013)
385e401.
[9] K. Ebrahimi, G.F. Jones, A.S. Fleischer, A review of data center cooling technology, operating conditions and the corresponding low-grade waste heat
recovery opportunities, Renew. Sustain. Energy Rev. 31 (2014) 622e638.
[10] S. Aghahosseini, I. Dincer, Comparative performance analysis of lowtemperature organic Rankine cycle (ORC) using pure and zeotropic working
uids, Appl. Therm. Eng. 54 (2013) 35e42.
[11] G. Manente, A. Toffolo, A. Lazzaretto, M. Paci, An organic Rankine cycle offdesign model for the search of the optimal control strategy, Energy 58
(2013) 97e106.
[12] M.O. Bamgbopa, E. Uzgoren, Quasi-dynamic model for an organic Rankine
cycle, Energy Convers. Manag. 72 (2013) 117e124.
[13] Y.R. Li, J.N. Wang, M.T. Du, S.Y. Wu, C. Liu, J.L. Xu, Effect of pinch point temperature difference on cost-effective performance of organic Rankine cycle,
Int. J. Energy Res. 39 (2013) 1952e1962.
[14] Y.R. Lee, C.R. Kuo, C.C. Wang, Transient response of a 50 kW organic Rankine
cycle system, Energy 48 (2012) 532e538.
[15] Y.R. Lee, C.R. Kuo, C.H. Liu, B.R. Fu, J.C. Hsieh, C.C. Wang, Dynamic response of a
50 kW organic Rankine cycle system in association with evaporators, Energies
7 (2014) 2436e2448.
[16] S.W. Hsu, H.W.D. Chiang, C.W. Yen, Experimental investigation of the performance of a hermetic screw-expander organic Rankine cycle, Energies 7
(2014) 6172e6185.
[17] B.R. Fu, S.W. Hsu, C.H. Liu, Y.C. Liu, Statistical analysis of patent data relating to
the organic Rankine cycle, Renew. Sustain. Energy Rev. 39 (2014) 986e994.

345

[18] B.R. Fu, S.W. Hsu, Y.R. Lee, J.C. Hsieh, C.M. Chang, C.H. Liu, Effect of off-design
heat source temperature on heat transfer characteristics and system performance of a 250-kW organic Rankine cycle system, Appl. Therm. Eng. 70
(2014) 7e12.
[19] B.R. Fu, S.W. Hsu, Y.R. Lee, J.C. Hsieh, C.M. Chang, C.H. Liu, Performance of a
250 kW organic Rankine cycle system for off-design heat source conditions,
Energies 7 (2014) 3684e3694.
[20] D. Manolakos, G. Papadakis, S. Kyritsis, K. Bouzianas, Experimental evaluation
of an autonomous low-temperature solar Rankine cycle system for reverse
osmosis desalination, Desalination 203 (2007) 366e374.
[21] D. Manolakos, G. Kosmadakis, S. Kyritsis, G. Papadakis, On site experimental
evaluation of a low-temperature solar organic Rankine cycle system for RO
desalination, Sol. Energy 83 (2009) 646e656.
[22] V. Lemort, S. Quoilin, C. Cuevas, J. Lebrun, Testing and modeling a scroll
expander integrated into an organic Rankine cycle, Appl. Therm. Eng. 29
(2009) 3094e3102.
[23] S. Quoilin, V. Lemort, J. Lebrun, Experimental study and modeling of an
organic Rankine cycle using scroll expander, Appl. Energy 87 (2010)
1260e1268.
[24] J.L. Wang, L. Zhao, X.D. Wang, A comparative study of pure and zeotropic
mixtures in low-temperature solar Rankine cycle, Appl. Energy 87 (2010)
3366e3373.
[25] X.D. Wang, L. Zhao, J.L. Wang, W.Z. Zhang, X.Z. Zhao, W. Wu, Performance
evaluation of a low-temperature solar Rankine cycle system utilizing R245fa,
Sol. Energy 84 (2010) 353e364.
[26] G. Pei, J. Li, Y. Li, J. Ji, Construction and dynamic test of a small scale organic
Rankine cycle, Energy 36 (2011) 3215e3223.
[27] G. Qiu, Y. Shao, J. Li, H. Liu, S.B. Riffat, Experimental investigation of a biomassred ORC-based micro-CHP for domestic applications, Fuel 96 (2012)
374e382.
[28] N. Zheng, L. Zhao, X.D. Wang, Y.T. Tan, Experimental verication of a rollingpiston expander that applied for low-temperature organic Rankine cycle,
Appl. Energy 112 (2013) 1265e1274.
[29] S. Declaye, S. Quoilin, L. Guillaume, V. Lemort, Experimental study on an opendrive scroll expander integrated into an ORC (organic Rankine cycle) system
with R245fa as working uid, Energy 55 (2013) 179e183.
[30] B. Twomey, P.A. Jacobs, H. Gurgenci, Dynamic performance estimation of
small-scale solar cogeneration with an organic Rankine cycle using a scroll
expander, Appl. Therm. Eng. 51 (2013) 1307e1316.
[31] M. Jradi, S. Riffat, Experimental investigation of a biomass-fuelled micro-scale
tri-generation system with an organic Rankine cycle and liquid desiccant
cooling unit, Energy 71 (2014) 80e93.
[32] V.K. Avadhanula, C.S. Lin, Empirical models for a screw expander based on
experimental data from organic Rankine cycle system testing, J. Eng. Gas
Turbines Power Trans. ASME 136 (2014) 062601.
[33] P. Klonowicz, A. Borsukiewicz-Gozdur, P. Hanausek, W. Kryowicz,
D. Brggemann, Design and performance measurements of an organic vapour
turbine, Appl. Therm. Eng. 63 (2014) 297e303.
[34] J.C. Chang, C.W. Chang, T.C. Hung, J.R. Lin, K.C. Huang, Experimental study and
CFD approach for scroll type expander used in low-temperature organic
Rankine cycle, Appl. Therm. Eng. 73 (2014) 1444e1452.
[35] Y.Q. Zhang, Y.T. Wu, G.D. Xia, C.F. Ma, W.N. Ji, S.W. Liu, K. Yang, F.B. Yang,
Development and experimental study on organic Rankine cycle system with
single-screw expander for waste heat recovery from exhaust of diesel engine,
Energy 77 (2014) 499e508.
[36] Z.Q. Wang, N.J. Zhou, J. Guo, X.Y. Wang, Fluid selection and parametric optimization of organic Rankine cycle using low temperature waste heat, Energy
40 (2012) 107e115.
[37] E.H. Wang, H.G. Zhang, B.Y. Fan, M.G. Ouyang, Y. Zhao, Q.H. Mu, Study of
working uid selection of organic Rankine cycle (ORC) for engine waste heat
recovery, Energy 36 (2011) 3406e3418.
[38] E.W. Lemmon, M.L. Huber, M.O. McLinden, NIST Standard Reference Database
23: Reference Fluid Thermodynamic and Transport PropertieseREFPROP,
Version 9.0, National Institute of Standards and Technology, Standard Reference Data Program, Gaithersburg, USA, 2010.
[39] L. Moroz, C.R. Kuo, O. Guriev, Y.C. Li, B. Frolov, Axial turbine ow path design
for an organic Rankine cycle using R-245fa, in: Proceedings of ASME Turbo
Expo 2013: Turbine Technical Conference and Exposition, Paper No.
GT2013e94078, pp. V05AT23A004, San Antonio, Texas, USA, June 3e7, 2013.
s, R. Collado, A. Mota-Babiloni, Performance
[40] B. Peris, J. Navarro-Esbr, F. Mole
evaluation of an organic Rankine cycle (ORC) for power applications from low
grade heat sources, Appl. Therm. Eng. 75 (2015) 763e769.
[41] R.B. Peterson, H. Wang, T. Herron, Performance of a small-scale regenerative
Rankine power cycle employing a scroll expander, Proc. Ins. Mech. Eng. Part A
J. Power Energy 222 (2008) 271e282.
[42] H. Wang, R.B. Peterson, T. Herron, Experimental performance of a compliant
scroll expander for an organic Rankine cycle, Proc. Ins. Mech. Eng. Part A J.
Power Energy 223 (2009) 863e872.
[43] J.A. Mathias, J.J.R. Johnston, J. Cao, D.K. Priedeman, R.N. Christensen, Experimental testing of gerotor and scroll expanders used in, and energetic and
exergetic modeling of, an organic Rankine cycle, J. Energy Resour. Technol.
Trans. ASME 131 (2009) 012201.
[44] V. Lemort, S. Declaye, S. Quoilin, Experimental characterization of a hermetic
scroll expander for use in a micro-scale Rankine cycle, Proc. Inst. Mech. Eng.
Part A J. Power Energy 226 (2012) 126e136.

346

B.-R. Fu et al. / Applied Thermal Engineering 80 (2015) 339e346

[45] R. Bracco, S. Clemente, D. Micheli, M. Reini, Experimental tests and modelization


of a domestic-scale ORC (organic Rankine cycle), Energy 58 (2013) 107e116.
[46] G. Liu, Y. Zhao, Q. Yang, L. Wang, B. Tang, L. Li, Theoretical and experimental
research on scroll expander used in small scale organic Rankine cycle System,
Proc. Inst. Mech. Eng. Part E J. Process Mech. Eng. 229 (2015) 25e35.
[47] N. Zhou, X. Wang, Z. Chen, Z. Wang, Experimental study on organic Rankine
cycle for waste heat recovery from low-temperature ue gas, Energy 55
(2013) 216e225.
[48] T. Saitoh, N. Yamada, S. Wakashima, Solar Rankine cycle system using scroll
expander, J. Environ. Eng. 2 (2007) 708e719.
[49] M.A. Tarique, I. Dincer, C. Zamrescu, Experimental investigation of a scroll
expander for an organic Rankine cycle, Int. J. Energy Res. 38 (2014) 1825e1834.
[50] T. Li, J. Zhu, W. Fu, K. Hu, Experimental comparison of R245fa and R245fa/
R601a for organic Rankine cycle using scroll expander, Int. J. Energy Res. 39
(2015) 202e214.
[51] Z. Miao, J. Xu, X. Yang, J. Zou, Operation and performance of a low temperature
organic Rankine cycle, Appl. Therm. Eng. 75 (2015) 1065e1075.
[52] P. Gao, L. Jiang, L.W. Wang, R.Z. Wang, F.P. Song, Simulation and experiments
on an ORC system with different scroll expanders based on energy and exergy
analysis, Appl. Therm. Eng. 75 (2015) 880e888.

[53] V.M. Nguyen, P.S. Doherty, S.B. Riffat, Development of a prototype low temperature Rankine cycle electricity generation system, Appl. Therm. Eng. 21
(2001) 169e181.
[54] H. Liu, G. Qiu, Y. Shao, F. Daminabo, S.B. Riffat, Preliminary experimental investigations of a biomass-red micro-scale CHP with organic Rankine cycle,
Int. J. Low-Carbon Technol. 5 (2010) 81e87.
[55] M. Li, J. Wang, L. Gao, X. Niu, Y. Dai, Performance evaluation of a turbine used
in a regenerative organic Rankine cycle, in: Proceedings of ASME Turbo Expo
2012: Turbine Technical Conference and Exposition, Paper No. GT2012-68441,
June 11e15, 2012, pp. 425e432. Copenhagen, Denmark.
[56] S.H. Kang, Design and experimental study of ORC (organic Rankine cycle) and
radial turbine using R245fa working uid, Energy 41 (2012) 514e524.
[57] M. Li, J. Wang, W. He, L. Gao, B. Wang, S. Ma, Y. Dai, Construction and preliminary test of a low-temperature regenerative organic Rankine cycle (ORC)
using R123, Renew. Energy 57 (2013) 216e222.
[58] A. Borsukiewicz-Gozdur, Experimental investigation of R227ea applied as
working uid in the ORC power plant with hermetic turbogenerator, Appl.
Therm. Eng. 56 (2013) 126e133.

Vous aimerez peut-être aussi