You are on page 1of 16

Energy Conversion and Management 144 (2017) 1833

Contents lists available at ScienceDirect

Energy Conversion and Management


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

An innovative small-scale two-stage axial turbine for low-temperature


organic Rankine cycle
Ayad M. Al Jubori a,b,, Raya Al-Dadah a, Saad Mahmoud a
a
The University of Birmingham, School of Engineering, Edgbaston, Birmingham B15 2TT, UK
b
University of Technology, Baghdad, Iraq

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

Article history: High expander efficiency is required to achieve best performance for small-scale organic Rankine cycle
Received 9 December 2016 (ORC) systems driven by low-temperature (<100 C) heat sources. In this paper a small-scale two-stage
Received in revised form 15 March 2017 axial turbine is modelled and compared with a single-stage axial turbine, with the aim of enhancing
Accepted 11 April 2017
the ORC performance by increasing its pressure ratio. The preliminary mean-line design approach is cou-
pled with three-dimensional CFD modelling and ORC cycle analysis was used to assess the impact of two-
stage axial turbine on the ORC cycle performance. Three-dimensional CFD analysis of the single and two-
Keywords:
stage axial turbines was performed using ANSYS17-CFX software. The RANS equations with a k-x SST
Organic Rankine cycle (ORC)
Small-scale
turbulence model were solved for three-dimensional viscous steady state flow. The real gas thermody-
Two-stage axial turbine namic properties of three organic working fluids (n-pentane, R141b, R245fa) are used in modelling the
CFD flow with both turbine configurations. Results revealed that the two-stage axial turbine configuration
Organic fluids exhibited a substantially higher turbine performance, with overall isentropic efficiency of 83.94% and
Low-temperature heat sources power output of 16.037 kW, compared to 78.30% and 11.06 kW from the single-stage configuration, with
n-pentane as working fluid and mean diameter of 64 mm for the two-stage configuration. Also, results
showed that the maximum ORC thermal efficiency was 14.19% compared with 10.5% for single-stage con-
figuration using n-pentane as the working fluid. These results highlight the potential of using two-stage
axial turbine in a small-scale ORCs system for the conversion of low-temperature heat sources into
electricity.
2017 Elsevier Ltd. All rights reserved.

1. Introduction appropriate for many electricity generation applications, such as


domestic and rural area and remote off-grid communities.
Due to the development in the world economy, as well as pop- The turbine is considered as a key component in the ORC sys-
ulation growth, the demand of energy consumption has directly tem and has substantial influence on cost, size and system perfor-
increased. This is based on fossil fuel consumption that has led to mance. Therefore, many studies have been performed extensively
serious environmental issues such as pollution, global warming on the expander to be used in the ORC system. The mean-line
and ozone depletion. The organic Rankine cycle (ORC) system is design of radial-inflow turbine (RIT) has been considered in many
one of the most effective technologies in harnessing low- studies; Sauret and Rowlands [1] performed mean-line design of
temperature heat sources (LTHS) like solar and geothermal energy an RIT for a geothermal heat-source driven ORC system with five
that cannot be used with conventional power systems. High den- organic working fluids. Their results showed that the power output
sity organic fluids are adopted in the Rankine cycle instead of varied from 338 kW with R134a to 254 kW with n-pentane; while
steam. The ORC offers distinctive advantages such as high reliabil- the turbine isentropic efficiency was 77%, which was approxi-
ity for low-medium grade heat sources, easy maintenance and low mately similar for all working fluids. Fiaschi et al. [2] built a
capital cost compared with the conventional Rankine cycle. Small- mean-line design model for an RIT for the ORC system. The perfor-
scale ORC systems based on single and two-stage axial turbines are mance of six organic working fluids was investigated for an output
power of 50 kW. The R134a had the maximum turbine isentropic
efficiency of 83% and inlet total pressure and temperature, rota-
tional speed and rotor inlet diameter of 38 bar, 147 C,
Corresponding author at: The University of Birmingham, School of Engineering,
54,000 rpm and 48 mm respectively. Pan and Wang [3] improved
Edgbaston, Birmingham B15 2TT, UK.
the ORC analysis by replacing the assumption of constant turbine
E-mail addresses: ama232@bham.ac.uk, ayadms@gmail.com (A.M. Al Jubori).

http://dx.doi.org/10.1016/j.enconman.2017.04.039
0196-8904/ 2017 Elsevier Ltd. All rights reserved.
A.M. Al Jubori et al. / Energy Conversion and Management 144 (2017) 1833 19

Nomenclature

A area (m2) Subscript/superscript


B axial chord (m) 17 station within the turbine and cycle respectively
C absolute velocity (m s1) accel accelerating
c chord length (m) AS aspect ratio
CL lift coefficient () blds blade
D, d diameter (m) c condenser
f correction/friction coefficients () cr critical
h specific enthalpy (kJ kg1) e evaporator
H blade height (m) ex exergy
K losses coefficient () H high
k specific turbulence kinetic energy (m2 s2) is isentropic
_
m mass flow rate (kg s1) hyd hydraulic
Nb number of blade () L low
ns specific speed () nbp normal boiling point
o throat (m) P profile/pump
p pressure (bar) R rotor
Q_ heat (kW) Re Reynolds number
Rn reaction () Rec recuperator
rm/rtip mean/tip radius (m) S stator
S blade space (pitch) (m) Sec secondary
s entropy (kJ kg1 K1) sh shock
T temperature (K) t turbine
t time (s)/blade thickness (m) T total
U blade velocity (m s1)/mean flow velocity (m s1) TC tip clearance
V velocity (m s1) TE trailing edge
W relative velocity (m s1) ts total-to-static
w specific work (kJ kg1) tt total-to-total
W _ power (kW) uncorrected

Greek symbols Acronyms


a absolute flow angle (degree) 1D, 3D one and three dimensional
b relative flow angle (degree) CFD computational fluid dynamics
g efficiency (%) GWP global warming potential
/ flow coefficient () LTHS low-temperature heat source
w loading coefficient () ODP ozone depletion potential
x specific turbulence dissipation rate (m2 sec3) ORC organic Rankine cycle
X rotational speed (rad s1) PD preliminary mean-line design
s tip clearance (m) RANS Reynolds-averaged Navier-Stokes
h stagger angle (degree) RIT radial-inflow turbine
f enthalpy loss coefficient () SST shear stress transport

efficiency with dynamic efficiency, based on the operating condi- Rahbar et al. [8,9] carried out an integrated mathematical approach
tions and working fluids. Conventional analysis based on the RIT for the development of an efficient small-scale RIT. In their inte-
was conducted. Their results showed that maximum dynamic effi- grated approach, the mean-line design was coupled with an opti-
ciency, power output and thermal efficiency were 75.75%, 8.25 kW mization technique and real gas formulation. Their results
and 9.0% respectively. Costall et al. [4] presented a detailed mean- showed that R152a has the highest turbine efficiency of 84%. White
line design methodology for an RIT. The design with a small power and Sayma [10] presented a 10 kWe small-scale ORC based on a
output such as 15.5 kW led to impractical blade geometry; while radial-inflow turbine for low-temperature applications like solar
the design of the medium power output of 34.1 kW achieved tur- energy and heat recovery. Their optimization results indicated that
bine isentropic efficiency of 51.50% at a rotational speed R1234ze was the optimum working fluid with a power output of
71,500 rpm. The third turbine design produced a power output of 7.32 kW corresponding with cycle efficiency of 7.26%. Li and Ren
45.6 kW with a turbine efficiency of 56.1%. Fiaschi et al. [5] pro- [11] carried out thermodynamic analysis of ORCs based on an RIT
posed a 1D model to design a 50 kW RIT for ORC applications with R123 as the working fluid. The turbine isentropic efficiency,
and the different methods of design were screened and analyzed system thermal efficiency and net power were 84.33%, 13.5% and
based on six organic working fluids. The design results showed that 534 kW respectively. Song et al. [12] investigated the effect of an
the range of total-to-total turbine isentropic efficiency was RITs performance on the analysis of an ORC system related to
between 72% and 80%, depending on the selected working fluid. operating conditions and working fluids.
Hu et al. [6,7] presented geothermal ORC performance analysis In terms of 3D CFD of RIT, Harinck et al. [13] performed a 3D
based on the design and off-design condition of an RIT. The work- CFD study on a Tri-O-Gen radial turbine for the ORC system oper-
ing fluid R245fa was considered with a turbine inlet temperature of ated by toluene. The 2D optimization was performed for a nozzle
65.9 C and a mass flow rate of 5.58 kg/s. Their maximum power which was manufactured and tested with 5 kWe power output.
output and thermal efficiency were 66.9 kW and 5.5% respectively. Sauret and Gu [14] presented the complete design and simulation
20 A.M. Al Jubori et al. / Energy Conversion and Management 144 (2017) 1833

process of a 400 kW radial-inflow turbine with R143a as the work- The R245fa showed a 10% higher power output compared with the
ing fluid in realistic geothermal conditions. The maximum turbine SES36. Pei et al. [27] constructed and dynamically tested a small-
isentropic efficiency was 83.5% with a mass flow rate and pressure scale ORC system based on a radial-inflow turbine (RIT) with
ratio of 17.24 kg/s and 2.64 respectively. Fiaschi et al. [15] carried R123 as a working fluid and the designed power output of
out 1D design and 3D CFD analysis of the rotor of a micro radial 3.3 kW. Their experimental results showed the turbine isentropic
turbine for a small, distributed ORC power system based on efficiency, cycle thermal efficiency and power output of 65%, 6.8%
R134a as a working fluid. Their comparison between 1D and 3D and 1.36 kW respectively. Ssebabi et al. [28] replaced the rotor of
showed that the maximum difference was 11.6% in terms of power a radial turbine kit with in-house manufactured and designed rotor
output. The CFD results indicated that the turbine total-to-static for low grade waste heat recovery. Their test was carried out using
efficiency and power output were of 71.76% and 5.162 kW at a air as a working fluid and then utilized to scale the turbine for
mass flow rate of 0.25 kg/s. The turbine efficiency with R123 R123. The experimental performance was very similar for both
reached 80.8% with a system thermal efficiency of 8.2% at a heat rotors with low isentropic efficiency (610%) compared to the pre-
source temperature of 393.15 K. Russell et al. [16] conducted the dicted efficiency of 13.7% with R123. Kang [29] designed an RIT
design and manufacture process of a 7 kW RIT using R245fa as using ideal gas equations and the development of a 30 kW ORC
the working fluid. The CFD performance was compared against a system with R245fa as the working fluid. The experimental and
mean-line design where the predicted turbine isentropic efficiency modelling results showed that the maximum average cycle and
by CFD was substantially lower by 69%. The CFD results indicated turbine efficiencies were found to be 5.22% and 78.8% respectively,
that maximum total-to-total efficiency was about 76%. Nithesh and with a maximum power output of 32.7 kW. Kang [30] designed a
Chatterjee [17] developed a small-scale 2 kW RIT for ocean thermal two-stage RIT to enhance the ORCs system performance working
energy conversion application based on R134a as a working fluid with fluid R245fa. The results showed that the turbine efficiency,
rotating at 22,000 rpm. Their results indicated that turbine effi- power output and thermal efficiency were 68.5%, 39.0 kW and
ciency was 70% with a rotor tip radius of 35.5 mm. 9.8% respectively at an evaporation temperature of 116 C and an
For an axial turbine, Moroz et al. [18] presented the detailed expansion ratio of 11.6. Cho et al. [31] conducted an experimental
design of a 250 kW axial turbine for an ORC power unit with study on a partial admission axial turbine for an ORC system with
R245fa in low temperatures up to 150 C. The structural optimiza- the rate of the partial admission of 16.7%. Based on their experi-
tion was carried out to reduce the rotor weight at an acceptable mental results, the thermal cycle efficiency was 2% at the total inlet
stress. The reported turbine efficiency achieved as a result from temperature of 100 C. Pu et al. [32] carried out an experimental
turbine optimization was 81.7%. Da Lio et al. [19,20] provided a study of a small-scale ORC system based on a single-stage axial
design criterion of an axial turbine to get a reliable estimate of tur- turbine with R245fa and HFE7100 as the working fluids. The max-
bine isentropic efficiency, which is obtained by mean-line design imum turbine efficiency, power and thermal efficiency were found
for a wide range of operating conditions and working fluids. The to be 59.7%, 1.979 kW and 4.01% respectively with a heat source
effect of working fluids was concluded on the efficiency in addition temperature less than 100 C. Sung et al. [33] developed and tested
to the expansion ratio and size parameter. While the 3D CFD anal- a design of a 200 kW ORC for waste heat recovery application
ysis of axial turbine has been conducted in [2123]; Cho et al. [21] based on an RIT; R245fa was used as a working fluid with a tem-
used a partial admission radial turbo-expander with a supersonic perature of 140 C. The experimental results exhibited that the
nozzle to convert available fluctuating waste thermal energy to ORCs thermal efficiency and power output were 9.6% and
produce a small sized power output with working fluid R245fa. 177.4 kW respectively.
The results showed that the thermal efficiency was about 4% with In order to achieve high ORC system performance in small-sized
6 nozzles, to produce 3.8 kW with an inlet total temperature at the power output applications from an LTHS requires the development
inlet of the turbine of 60 C. Al Jubori et al. [22] presented a new of high turbine performance. It is evident there has been limited
methodology that integrates the ORC based on a small-scale axial attention given to using an axial turbine in a small-scale ORC for
turbine with mean-line design and 3D CFD analysis. Their results LTHS applications. Moreover, there is limited published literature
indicated that using working fluid R123 with the turbine mean on using axial turbines for small-scale ORC applications. Also, no
diameter of 70 mm, the maximum total isentropic efficiency of published studies regarding small-scale two-stage axial turbines
82% was achieved with a power output 5.6 kW. Al Jubori et al. with low mass flow rate and low-grade temperature heat sources
[23] developed micro-scale axial and radial-inflow turbines for a (<100 C) with different organic fluids. Therefore, this study aims
low-temperature heat source driven ORC based on five organic flu- to develop a novel small-scale two-stage axial turbine for an ORC
ids; 1D mean-line design and 3D CFD analysis were conducted. The system using an integrated ORC modelling with mean-line design
results showed that the axial turbine was competitive to the radial of an axial turbine using engineering equation solver (EES) soft-
turbine at a mass flow rate of 0.5 kg/s with a maximum ORC of ware and 3D CFD simulation using ANSYS17-CFX. To achieve an
10.60%, based on the radial turbine compared with 10.14% based accurate prediction and the real behavior of organic working fluids,
on the axial turbine. While Clemente et al. [24] assessed the perfor- real gas formulation is utilized in the ORC/turbines model. Finally,
mance of different expanders including axial and radial turbines the results inclusion design and off-design conditions are pre-
and scroll and positive displacement expanders based on 1D anal- sented and compared to a single-stage configuration in available
ysis, to design a bottoming cycle to recover heat from exhaust low-grade heat source such as solar and geothermal energy.
gases of a 100 kW gas turbine. The highest power from turbine
was 26 kW with 8% cycle efficiency.
For experimental investigation, Chang et al. [25] conducted an 2. ORC system layout and working fluids
experimental study of a low-temperature organic Rankine cycle
based on a scroll expander with R245fa as a working fluid; the Any recuperative ORC system consists of five classes of compo-
temperature of the heat source was below 100 C. The results indi- nents which are the evaporator, turbine, recuperator, condenser
cated that the expanders and the cycles efficiencies and power and pump as shown in Fig. 1. Other components besides them such
output were 73.1%, 9.44% and 2.3 kW respectively. Ziviani et al. as a generator and system operation and control are generally
[26] performed a numerical and experimental study of a single- required. The organic fluid in a liquid state is pressurized and deliv-
screw expander for ORC applications with a heat source tempera- ered to the evaporator by the pump. In the evaporator, the liquid is
ture of 125 C; R245fa and SES36 were used as the working fluids. vapourized and then the vapour is passed to the turbine where it
A.M. Al Jubori et al. / Energy Conversion and Management 144 (2017) 1833 21

expands to generate power. The superheated vapour leaves the The second law efficiency (exergy efficiency) can be defined as
turbine at high temperature; the recuperator can be used to the proximity of the real thermal efficiency of the cycle to the Car-
recover the thermal energy from the fluid at the turbine exit. After not cycle efficiency as:
the recuperator, the vapour is condensed to liquid through the con-
denser. In this study, the subcritical ORC is investigated to avoid gth W_
gex  net  9
high-pressure system complexity and safety concerns. The heat gCarnot Q_ e 1  TL
TH
and pressure losses through the connecting pipe of the ORC system
are neglected. The assumption of a steady state operating condi- According to ORC cycle modelling, there are several parameters
tion is applied on the system. The power consumption by the which need to be specified as input/design parameters. These
pump is given by: parameters in terms of inlet total temperature (heat source tem-
_ perature) and outlet temperature (cold side temperature) with
_ p mh6s  h5
W 1 three working fluids for different mass flow rates are listed in
gp
Table 1. Also, the pump, generator and mechanical efficiencies
The heat added from the LTHS is given by: and recuperator effectiveness were selected based on the litera-
ture. The working fluid mass flow rate is used as one of the input
_ e mh
Q _ 1  h7 2 parameters in ORC system modelling to calculate the desired
The power output from the turbine is given by: power output. Therefore, the turbine can be sized to meet ORC sys-
tem specification and thus the turbine and ORC system perfor-
_ t m
W _ h1  h2s gt1 m
_ h2  h3s gt2 3 mance can be expressed as a function of mass flow rate.
The heat rejected through the condenser is given by: The organic working fluids for low-grade temperature heat
source applications have a low-boiling temperature. Many low-
_ c m
Q _ h4  h5 4 boiling organic fluids such as chlorofluorocarbon (CFC), hydrochlo-
rofluorocarbon (HCFC) and hydrofluorocarbon (HFC) are offered on
The heat transfer between the superheated vapour and the
the market. The thermo-physical properties have an influence on
under-cooled fluid at the pump exit through the recuperator is
the architecture and cost, and on the components size of an ORC
given by the following equation:
system. The thermo-physical properties, environmental impact
_ Rec m
Q _ h3  h4 m
_ h7  h6 5 and safety aspects differ from one working fluid to another. Also,
the working fluids should be environmentally friendly with low
The effectiveness is the ratio between the effectively transferred
global warming potential (GWP) and zero ozone depletion poten-
heat load and the maximum heat load that could be transferred in
tial (ODP) and have low flammability and corrosion characteristics
an ideal regenerator and it is given by the following equation:
as clarified in Table 2.
T7  T6
ERec 6
T3  T6
where ERec is the recuperator effectiveness. Table 1
The net power output from the ORC cycle is given by: The input parameters of ORC model.

_ net W
_ tg _ Parameters Unit Value
W mech ggen  Wp 7
Heat source temperature K 365
where the gmech and ggen are mechanical efficiency and generator Heat sink temperature K 298
efficiency. Pump efficiency 0.75
Generator efficiency 0.96
The ORC thermal efficiency is given by: Mechanical efficiency 0.96
_
W
Recuperator effectiveness 0.8
gth _ net 8 Working fluid mass flow rate kg/s 0.20.5
Qe

Fig. 1. Schematic diagram of recuperative organic Rankine cycle.


22 A.M. Al Jubori et al. / Energy Conversion and Management 144 (2017) 1833

Table 2
Physical, safety, and environmental properties for three organic fluids.

Fluid Mol. weight (g/mol) Tnbp (K) Tcr (K) Pcr (kPa) ODP GWP (100 yr) Atmospheric life time (yr)
n-Pentane 72.15 309.1 469 3360 0 20 0.01
R141b 116.95 305.05 480 4460 0.12 725 9.2
R245fa 134.05 288.14 426 3610 0 950 7.7

3. Mean-line design of small-scale two-stage axial turbine The velocities across the turbine stage are calculated as:

C 2  2 9
The performance of an ORC system is considerably influenced 1  Rn W2 /2 >
1
>
>
by turbine efficiency and pressure ratio. Typically, the maximum
U
C 2   >
>
W 2 2>
=
pressure ratio of a single-stage turbine is identified to be around U
2
1  R n  2
/
W 2 2 W 2 11
4 [34]. Therefore, the design of a two-stage turbine is to compen-
2  Rn /2 >
>
>
>
sate for the weakness of the single-stage turbine in a high pressure U
W 3 2 W 2 >
>
2 ;
operating condition. The preliminary mean-line design (PD) of an U
2
R n /
axial turbine mostly depends on a mean-line flow model. Also,
the PD model is known as a 1D model that assumes unidirectional The expressions of the degree of reaction, flow and loading coef-
and uniform flow at the mid-span of the blade passages. The PD is ficient are as following:
formulated based on inlet and exit conditions of each blade pas- 9
sage (cascade) neglecting the blade shape. The PD is the most Rn DDhhstage
rotor
>
>
>
=
essential step of the whole initial design process.
CUx 12
The output from the PD provides the turbine layout in terms of >
>
velocity triangles, size, blade shape and height. The flow angles are W Uw2 rDhstage >
;
X2tip
calculated based on three dimensionless parameters which are
reaction, flow and loading coefficients as detailed in Eqs. (10) and The specific speed is a dimensionless parameter and defined by
(11); where the reaction is defined for turbomachinery as the static Eq. (13) as stated in [34,37,38] as follows:
enthalpy drop through the turbine rotor to the static enthalpy drop
q
through the turbine stage (stator and rotor). Loading coefficient X _
m
qexit
expresses the measure of the work done (work out from turbine ns 0:75
13
stage) by turbine stage or the drop in pressure/temperature Dhis
through the turbine. The flow coefficient is the third dimensionless
parameters which reflects the effect of the mass flow rate (axial Based on the system of units (i.e. SI for Dixon [38] and imperial
velocity) and mean velocity of the blade turbine. for Balje [34]), which deliver different specific speed values with
Fig. 2 shows the velocity triangle for the axial turbine stage. In a the same characteristic for turbines configurations. The specific
single-stage axial turbine, a1 = 0; while, a1 = a3 in a multi-stage speed was used with turbine efficiency to establish chart that
axial turbine. Thus, the flow angles at the inlet and outlet of the can be useful for suitable selection of turbine type based on the
turbine stage are defined as [3537]: available conditions. According to Dixon and Hall [38], specific
speed should lie between 0.3 and 1.5 to achieve the maximum
9
tan b2 W2R n
>
>
axial turbine design. Various operating point parameters at inlet/
2/
>
> outlet boundary conditions such as mass flow rate, densities, pres-
>
=
tan b3 W2/
2Rn
sure, and temperature are used to obtain the specific speed based
10
n >
tan a3 W=21R >
> on Eq. (13), where the rotational speed was limited to 30,000 rpm
/ >
>
W=21Rn ; in this equation. The calculated specific speed range using Dixons
tan a2 / and Baljes charts [34,38] was listed in Table 3.

C1
o (throat width)
1 2
S 2 W2 Rotor W3 U
Stator (Blade velocity)
B C2 3
C3
U 3
Axial chord

Fig. 2. Axial turbine blade passage.


A.M. Al Jubori et al. / Energy Conversion and Management 144 (2017) 1833 23

Table 3 a1b a1b   tmax =c a1b =a2


Input parameters of PD for both single and two-stage axial turbines and their K P K Pa1b 0 K Pa1b a2  K Pa1b 0
ranges/values.
a2 a2 0:2
21
Parameters Unit Values/Range
Loading coefficient (W) 0.61.2

2
Flow coefficient (/) 0.20.8 K P 0:914 K P K accel K sh 22
Reaction (Rn) 0.40.6 3
Hub/tip radius ratio (rh/rt) 0.60.75
Inlet total temperature K 365 The secondary loss occurs because of the vortices due to the
Degree of superheating K 010 boundary layer and the blade passage curvature which leads to
Inlet-total-pressure bar Corresponding saturated vapour the deviation of the working fluid from the main direction and it
pressure at inlet temperature is calculated according to the following equation [37]:
Condensing pressure bar Corresponding saturated pressure


2
cos a2 cos2 a2
at heat sink temperature
CL
Mass flow rate kg/s 0.20.5 KSec 0:0334f AS 23
Specific speed 0.5200.710 cos a1b S=c cos3 am
Rotational speed rpm 30,000
Working fluids n-pentane, R141b, R245fa The trailing edge loss coefficient is defined in terms of the ratio
between the thickness of the trailing edge and the throat of the
blade passage according to the following expression [19]:
The mean radius is the midway between the tip and hub and it

can be defined as the diameter that divides the annulus into two Dp0 t2
KTE 24
equal areas as follows [38]: 0:5qC2
2 o2  t2
 
r2m r2t rh =2
2
14 The loss coefficient of the blade clearance is calculated for the
The blade height can be determined from the following equa- rotor only as the following [37]:
tion [38]:  s  cos2 a
tan a1  tan a2
2
KTC 4B 25
_
m H cosam
H rt  rh  15
2pr m qU The enthalpy loss coefficient can be defined as a measure of the
To obtain the turbine dimensions and blade shape including lost kinetic energy relative to exit kinetic energy. Where, the pres-
blade height and pitch, chord length, number of blades and leading sure loss coefficient and enthalpy loss coefficient are approxi-
and trailing edge thickness are detailed in [3538]. The blade pitch mately equal at a small value of enthalpy. [36] The full details of
and chord length can be calculated using Zweifels correlation and conversion from the pressure loss to enthalpy loss are outlined in
stagger angle as follow [38]: Moustapha et al. [37] The stage total-to-total and total-to-static

isentropic efficiencies in terms of enthalpy loss are as follows [38]:
S  
0:4 2cos2 tan a2 tan a1 16 1
B gtt h   i 26
1 fR W23 =2 fS C22 =2 h3 =h2 =h01  h03
S S
cos h 17
c B
1
where h is the stagger angle which can be calculated using the fol- gts h   i 27
1 fR W23 =2 fS C22 =2 h3 =h2 C23 =2 =h01  h03
lowing equation [38]:
1 The flow and loading coefficients are initially selected using
h a1 a2 18 Smith chart as detailed in Moustapha [37] and Dixon and Hall
2
[38] in order to have realistic turbine efficiency and velocity trian-
For rotor calculations, a angle should be replaced by b angles.
gle. The values of flow and loading coefficients and degree of reac-
The number of blade can be calculated by the following equa-
tion in Table 3 are within the optimal range of the axial turbine
tion [36,38]:
performance as reported in the literature based on the perfor-
2pr m mance correlation charts [3638]. As shown in Fig. 3, the PD is
Nb 19
S highly iterative procedure of turbine design and hence comprehen-
From aerodynamic point of view, it is preferable to make the sive studies are required for different configurations based on var-
trailing edge thickness as thin as possible, and the typical values ious input parameters that include reaction, flow and loading
between (0.0150.05c) as reported in [35]. For tip clearance, the coefficients. The flow chart of the PD procedure is shown in
typical limit of the manufacture is 0.35 mm as a minimum value. Fig. 3; the engineering equation solver (EES) [39] is used to develop
The losses model of AMDCKO (Ainley and Mathieson; Dunham the PD code which has the ability to deliver the initial blade shape
and Came; Kacker and Okapuu) is employed to develop the PD and the geometry dimensions of the axial turbine stage. The PD
code to calculate the losses within the blade passage. The AMDCKO code is coupled with ORC modelling (Eqs. (1)(9)). The different
losses model is used to predict the performance of a two-stage input parameters for the PD code are outlined in Table 3. Then
axial turbine. The total pressure loss coefficient (K) is defined as Table 4 provides the initial base-line geometry from the PD which
a measure of overall aerodynamics losses (stagnation pressure will be exported into ANSYS17-BladeGen to generate the three-
loss) through the blade row and can be subdivided into four com- dimensional blade geometry.
ponents of losses (i.e. profile, secondary, trailing edge and tip clear-
ance losses) and is expressed in the following equation as [36,37]: 4. CFD methodology
K T K P f Re K Sec K TE K TC 20
In turbine design, CFD analysis is the crucial technique, hand-
The profile loss coefficient based on the AMDCKO model is in-hand with the PD, to calculate the accurate turbine performance
given as [36,37]: for various working fluids and operating conditions of small-scale
24 A.M. Al Jubori et al. / Energy Conversion and Management 144 (2017) 1833

five steps, which are blade generation, grid generation, CFD set-
Input data
up, CFD solution and post-processing of the simulation analysis.
(Table 3)
After the preliminary mean-line design has been completed, the
principle turbine dimensions and blade shape for the turbine stage
Guess turbine efficiency based on (stator and rotor) are imported to the ANSYS17-BladeGen tool as a
and module of blade design, to generate the blade geometry for both
the stator and rotor for both turbine configurations, as shown in
Update turbine efficiency

Fig. 4. The pressure/suction and angle/thickness modes are


Determination of losses employed respectively, which allow the defining of the curves for
(section 3) the hub, shroud and blade profile for the stator and rotor blades.
The computational mesh is generated for the fluid domain
1D geometry calculation across the stator and rotor blade passages using ANSYS17-Turbo
(stator, rotor). Grid via hexahedral mesh; its mostly based on O-H grid. To allow
the Turbo-Grid to determine an appropriate topology for the blade
passage according to the blade angle, the leading edge and the
trailing edge type, ATM Optimized (Automatic topology and mesh-
Matching constraints ing) has been applied. A grid independence study is conducted for
No If each working fluid and turbine configuration to ensure that 3D CFD
|(ts)new-(ts)old|<10-3 analysis results are grid independent.
In 3D CFD analysis, RANS equations are solved with a k-x SST
(shear stress transport) turbulence model using the high resolution
Yes advection scheme. The k-x SST has the advantage of the ability to
capture the turbulence closure using the automatic near wall treat-
Final turbine geometry ment, by determining the dimensionless (y+) for the first node after
the wall. The dimensionless distance (y+) is maintained equal or
less than unity as recommended in the CFX users manual. The tur-
ANSYS BladeGen
bulence intensity at the inlets boundary condition is kept at 5%,
(blade generation)
which is the recommended value when no information about the
inlets turbulence quantities is available. The k-x transport equa-
ANSYS TurboGrid tions carried out to achieve the turbulent kinetic energy and the
(3D mesh generation) specific dissipation rate are:

@ @ @ @k
qk qkui Ck Gk  Yk SK 28
Turbine Assembly & 3D CFD @t @xi @xj @xj
Analysis via ANSYS CFX

@ @ @ @x
qx qxui Ck Gx  Yx Sx 29
@t @xi @xj @xj
Performance analysis for turbine and
ORC cycle where Gk and Gx represent the generation of turbulent kinetic
energy and its dissipation rate; Yk and Yx represent the fluctuating
Output dilation in compressible turbulence; Sk and Sx are the source terms
of the k-x turbulence model.
Fig. 3. Flow chart of PD and 3D CFD methodology implemented. A stage (mixing-plane) interface was specified at the interface
between the stator and the rotor to deliver a connection across
the domain of the stationary and rotating blade rows, to take into
single and two-stage axial turbines. Three-dimensional CFD analy- account that the flow at the interface is non-axisymmetric. The GGI
sis is conducted across the stator and rotor blade passage for both (generalized grid interface) is applied between any neighbouring
configurations using ANSYS software. The CFD analysis includes stationary parts. The CFD analysis was conducted at the nominal

Table 4
_ 0:5 kg=s and three investigated working fluids.
PD output of two-stage axial turbine for m

Parameters R245fa R141b n-Pentane


1st stage 2nd stage 1st stage 2nd stage 1st stage 2nd stage
Tip diameter (dt) mm 72 72 66 66 76 76
Hub diameter (dh) mm 58 58 54 54 60 60
Blade height (H) mm 7 7 6 6 8 8
Tip clearance (mm) 0.35 0.35 0.35 0.35 0.35 0.35
Stagger angle (deg) 33.0 28.50 35.0 31.75 30.25 27.50
LE blade angle (deg) 11 13 10 12 13 15
TE blade angle (deg) 64.75 68.25 63.75 66.50 68.50 70.25
Rotor number of blade () 24 20 22 18 26 22
Stator number of blade () 25 25 23 23 27 27
Solidity (c/S) () 1.712 1.675 1.527 1.642 1.763 1.840
Flow coefficient () 0.564 0.581 0.596 0.603 0.627 0.634
Loading coefficient () 0.865 0.894 0.962 1.034 0.981 1.125
Degree of reaction () 0.472 0.478 0.457 0.464 0.486 0.495
Turbine isentropic efficiency (%) 80.19 82.59 82.63 85.11 83.48 86.04
A.M. Al Jubori et al. / Energy Conversion and Management 144 (2017) 1833 25

Fig. 5. The grid sensitivity based on the efficiency of two-stage axial turbine with n-
pentane as a working fluid.

Fig. 4. 3D view of two-stage axial turbine geometry. with three organic working fluids, the maximum difference
between the CFD and PD in terms of efficiency was 4.17% with
R245fa as a working fluid; while the maximum difference in terms
design condition (Table 3) where the inlets total pressure and tem-
of power output was 3.38% with R245fa. The variance between CFD
perature were defined at the inlet with static pressure at the outlet.
and PD is mainly because of the characteristic of the 1D mean-line
In all 3D CFD simulations, the whole volume of the axial turbine
design which is not able to capture all properties of three-
stage is considered with a steady-state condition, turbulent viscous
dimensional flow fields. Also, it could be attributed to the complex
flow, single-phase and compressible flow. All the walls are
phenomena that can be accurately handled by 3D CFD modelling.
assumed to be adiabatic, non-slip and smooth. The convergence
Furthermore, the 3D CFD results of the two-stage turbine are com-
criterion of CFD analysis for all the residuals (RMS) through the
pared with Ref. [40]; most of the range/values of dimensions and
CFX was equal to 105. The grid independence study was con-
operating conditions are taken from the above-mentioned refer-
ducted for the turbine stage (stator and rotor) to confirm that the
ence as shown in Table 5 to allow a fair comparison. The compar-
grid size is sufficient and the CFD results are independent of the
ison results showed that the maximum difference in efficiency was
number of grid nodes. The initial computational grid of the 3D
about 3.57%, as shown in Fig. 8. This difference could also be attrib-
CFD simulation was generated and the solution found; the turbine
uted to the complex phenomena that can be accurately handled by
efficiency and dimensionless wall distance (y+) were calculated.
3D CFD modelling.
The computational mesh is then refined by adding more nodes in
both directions of hub to tip and blade-to-blade; the CFD simula-
tion was repeated and re-run until the solution of grid indepen- 6. CFD results
dence was reached, as shown in Fig. 5 with n-pentane as a
working fluid. The turbine performance in terms of efficiency and power out-
Fig. 5 displays the grid sensitivity analysis in terms of the tur- put is sensitive to the working fluids, expansion ratio, mass flow
bine efficiency for both configurations, where it shows that the tur- rate, rotational speed and turbine size at nominal design point
bine efficiency becomes stable and independent on the node count (Table 3) and off-design conditions as shown in Figs. 916 as para-
over 20,000,000; the same procedure is repeated for other working metric studies. Three organic working fluids are investigated in
fluids. The mesh density of the blade-to-blade passage for both order to explore which produces the best turbine and system per-
configurations is displayed in Fig. 6. The real fluid properties data- formance under available operating conditions. For each stage and
base in ANSYS17-CFX is integrated with REFPROP software, to working fluid, the 3D CFD analysis was conducted with the design
obtain an accurate thermodynamic model of the organic fluids operating conditions as displayed in Table 3. It is evident from
properties in the 3D CFD analysis of ORC turbines. The 3D CFD sim- Fig. 9 that the two-stage configuration has considerably higher
ulations were performed for a flow across the turbine stage using overall isentropic efficiency and power output at an expansion
an Intel CPU core i7 4820 K@ 3.70 GHz with 48 GB RAM mem- ratio of about 5.0 for all working fluids. Where the expansion ratio
ory in parallel run with 4 CPU cores. is defined based on the pressure ratio and the maximum expansion
ratio was around 5.0. It is also clear from Fig. 9 that the maximum
overall efficiency of the two-stage configuration was 83.94% com-
5. CFD verification and validation pared with 78.30% for the single-stage configuration with n-
pentane as a working fluid. Also, the reduction in turbine efficiency
The field of small-scale ORC systems generally suffers from a is due to the possibility of chocking occurs at large expansion ratio
lack of experimental data suitable for model validation. Therefore when the turbine stage runs at off-design points. It is shown in
verification of the present 3D CFD analysis results of both axial tur- Fig. 10 that the maximum total power output is significantly higher
bines configurations was compared against the preliminary mean- in the two-stage configuration; where the maximum delivered
line design (PD) results at the operating conditions (Table 3) in power output from the two-stage configuration was 16.037 kW,
terms of turbine performance (isentropic efficiency and power out- compared with 11.06 kW from the single-stage configuration with
put), as illustrated in Fig. 7. For both axial turbine configurations n-pentane as the working fluid. The R245fa delivered the minimum
26 A.M. Al Jubori et al. / Energy Conversion and Management 144 (2017) 1833

Fig. 6. Computational grid for two-stage axial turbine configuration.

Fig. 7. Comparison between preliminary mean-line design (PD) and CFD for two-stage axial turbine.

Table 5
Input parameters and two-stage geometry Ref. [40].

Input parameters Turbine geometry


Property Value Property Value
Working fluid () Argon Rotor number 36
of blades ()
Inlet total pressure (bar) 1.5826 bar Stator number 44, 40
of blades ()
Inlet total temperature (K) 936.11 K Tip diameter (mm) 246, 249
Mass flow rate (kg/s) 0.277 Mean diameter 215.9
(mm)
Total-to-total pressure 1.226 Tip clearance (mm) 0.3, 0.38
ratio ()
Total-to-static pressure 1.232 Inlet flow angle 12
ratio () (degree)
Flow coefficient () 0.465 Stator solidity () 1.57,
0.651 1.56
Rotational speed (rpm) 12,000 rpm Rotor solidity () 1.3, 1.4

power output for both turbines configurations. Lighter organic flu- Fig. 8. Comparison of the 3D CFD results from current model with Ref. [40] for two-
ids (low molecular weight such as n-pentane can produce higher stage axial turbine.
power outputs compared with heavier fluids (fluids with high
molecular weight/high density, such as R245fa for the same work- organic fluids have a higher enthalpy drops and consequently lar-
ing fluid mass flow rate, due to a relatively larger turbine size and ger specific work. The high-density organic fluids such as R245fa
consequently higher specific work output, whereas the lighter have smaller size due to their lower specific volumes and thus
A.M. Al Jubori et al. / Energy Conversion and Management 144 (2017) 1833 27

a) Singlestage b) Twostage

_ 0:5 kg=s and three working fluids.


Fig. 9. Variation of the total-to-static efficiency with expansion ratio for both turbine configurations at m

a) Singlestage b) Twostage

_ 0:5 kg=s and three working fluids.


Fig. 10. Variation of the power output with expansion ratio for both turbine configurations at m

a) Singlestage b) Twostage

Fig. 11. Variation of the total-to-static efficiency with mass flow rate for both turbine configurations and three working fluids.
28 A.M. Al Jubori et al. / Energy Conversion and Management 144 (2017) 1833

a) Singlestage b) Twostage

Fig. 12. Variation of the power output with mass flow rate for both turbine configurations and three working fluids.

a) Singlestage b) Twostage

Fig. 13. Variation of the total-to-static efficiency with rotational speed for both turbines configurations and three working fluids.

a) Singlestage b) Twostage

Fig. 14. Variation of the power output with rotational speed for both turbines configurations and three working fluids.
A.M. Al Jubori et al. / Energy Conversion and Management 144 (2017) 1833 29

a) Singlestage b) Twostage

Fig. 15. Variation of the total-to-static efficiency with turbine mean diameter for both turbine configurations and three working fluids.

a) Singlestage b) Twostage

Fig. 16. Variation of the power output with turbine mean diameter for both turbine configurations and three working fluids.

require a higher pressure ratio and rotational speed to achieve the has the maximum turbine isentropic efficiency and power output
same power output compared to the lighter fluid. for both turbines configurations with various mass flow rates.
Figs. 11a, b and 12a, b present the variation of the overall isen- It is evident from Figs. 13a, b and 14a, b that there exists a
tropic efficiency of the turbine and power output with the mass design condition (Table 3) for both configurations that harvests
flow rate for both turbines configurations respectively at nominal the highest overall isentropic efficiency and power output at the
operating conditions (i.e. Table 3). As can be seen from rotational speed of 30,000 rpm as a design condition. Moreover,
Fig. 11a and b that overall total-to-static isentropic efficiency these figures demonstrate that both overall isentropic efficiency
increases with the mass flow rate until the mass flow rate reaches and power output are a strong function of rotational speed as seen
of 0.5 kg/s as a design point in both turbines configurations. While when increasing the rotational speed from 20,000 rpm to
the overall turbines efficiency decreases (above the mass flow rate 40,000 rpm as off-design conditions with both configurations and
of 0.5 kg/s as the off-design condition) because of the potential of three different working fluids. The maximum overall isentropic
chocking arises at large mass flow rate. It is found from efficiency was 83.94% for the two-stage compared with 78.30%
Fig. 12a and b that power output increases as the mass flow rate for the single-stage configuration with n-pentane as a working
increases for both configurations and three different working flu- fluid. As one can note from Figs. 13a, b and 14a, b with increasing
ids. This is associated with the definition of power output as the the rotational speed, the turbine performance (efficiency and
relation between the mass flow rate and the enthalpy drop. With power) enhanced and reached the optimum values of power and
the rise of working fluid mass flow rate, the actual enthalpy drop efficiency at the design point. The reduction in turbine perfor-
increases leading to larger power output (as seen in mance (i.e. efficiency and power) shows the influence of off-
Fig. 12a and b). As depicted in Fig. 11a and b the minimum turbine design operation on turbine performance. The R245fa showed the
efficiency was 69.50% for the single-stage configuration, compared minimum performance for both turbines configurations. The opti-
with about 73.0% for the two-stage configuration at a mass flow mum specific speeds of the three working fluids namely n-pentane,
rate of 0.2 kg/s with R245fa as a working fluid. While n-pentane R141b and R245fa were 0.707, 0.587 and 0.525 respectively.
30 A.M. Al Jubori et al. / Energy Conversion and Management 144 (2017) 1833

The effect of turbine mean diameter on turbine performance


(efficiency and power) for both configurations and three working
fluids is substantial, as depicted in Figs. 15a, b and 16a, b. With
the rise of the mean diameter based on the increase in the mass
flow rate within design range 0.20.5 kg/s, the turbine efficiency
and the actual enthalpy drop increase leading to a larger power
output based on the definition of the loading coefficient. The best
performance for the two-stage configuration was at a mean diam-
eter of about 64.0 mm compared with about 70 mm for the single-
stage configuration at mass flow rate of 0.5 kg/s and n-pentane as
the working fluid. Based on Eqs. (14) and (15), a larger blade height
(i.e. the increase of the turbine size as shown in Figs. 15 and 16) is
achieved at design point which leads to a reduction in the blade tip
clearance losses, and possibly, other losses (e.g. secondary loss).
When the mean diameter increases over 64 mm for n-pentane,
61 mm for R141b and 57.5 mm for R245fa as the design points
for two-stage configuration, the overall turbine efficiency
decreased because of the turbine works at off-design condition.
The same behavior can be observed with single-stage configuration
when the mean diameter increases over 70 mm for n-pentane, Fig. 18. Blade loading for 1st and 2nd stage of axial turbine configuration with n-
65.5 mm for R141b and 61.25 mm for R245fa as the design points. pentane as the working fluid.
It is found from Fig. 17a, b for both turbines configurations,
lighter organic fluids (low molecular weight of working fluids as
in n-pentane) can yield a significantly larger turbine mean diameter
compared with a heavier organic fluid such as R245fa, due to lower
specific volume. Depending on the loading coefficient definition,
increasing mass flow rate increases the actual enthalpy drop lead-
ing to a larger turbine mean diameter; also based on the relation-
ship between the mass flow rate and turbine diameter (i.e. Eq. (15)).
The pressure distribution of the rotor passage for the two-stage
configuration is revealed in Fig. 18; it indicates the highest values of
the pressure on the pressure side, corresponding to the location of
the low velocity of the flow as shown in velocity vector in Fig. 19
due to the high profile curvature in this zone. In contrast, the lowest
values of pressure are located on the suction side, due to the highest
values of the flow velocity at the throat area of the blade passage as
can be seen in Fig. 19. The isentropic enthalpy drop (specific work)
is provided by the area circumscribed by such pressure distribution
curves; where, the enclosed area is indicative of the net torque pro-
ducing aerodynamic force by the rotor turbine shaft. Fig. 20 shows
the contour of pressure at 50% of the span. It is noticeable that the
maximum pressure is at the inlet of the stator of the first stage, as Fig. 19. Velocity vector for the 1st stage of dual stage axial turbine configuration
shown in Fig. 19 for three organic working fluids. with n-pentane as the working fluid.

a) Singlestage b) Twostage

Fig. 17. Variation of turbine mean diameter with mass flow rate for both turbine configurations and three working fluids.
A.M. Al Jubori et al. / Energy Conversion and Management 144 (2017) 1833 31

c) R245fa
a) n-pentane b) R141b

Fig. 20. Pressure distribution contour for two-stage axial turbine configuration for three working fluids.

7. ORC analysis results

The turbine performance parameters such as isentropic effi-


ciency and power output, obtained from the 3D CFD analysis for
each working fluid, were then inserted as the inputs into the ORCs
model (Eqs. (1)(9)) to calculate the ORCs thermal efficiency at
nominal operating conditions (i.e. Table 1) as shown in Fig. 21.
Where the turbine performance (efficiency and power) is dynami-
cally obtained based on PD model and then by 3D CFD analysis
with thermodynamic operating conditions for each working fluid.
It is evident that increasing the mass flow rate of working fluids
leads to an increase in the ORC systems efficiency. In both config-
urations, the power output increased with the rise in the mass flow
rate, which led to higher ORC thermal efficiency based on its defi-
nition. As can be seen from Fig. 21, the maximum ORC efficiency
was about 14.19% for the two-stage configuration, compared with
10.50% for the single-stage configuration with n-pentane as a
working fluid. Moreover, it can be found from Fig. 22a at nominal
operating conditions (Table 3), R245fa has the minimum ORC ther-
Fig. 21. ORC thermal efficiency with different mass flow rates for three organic
mal efficiency in both two-stage and single-stage configurations, at
fluids and both configurations.

Fig. 22. Cycle thermal efficiency (a); and second law efficiency (b); for both turbines configurations with three working fluids.
32 A.M. Al Jubori et al. / Energy Conversion and Management 144 (2017) 1833

about 12.03% and 8.79% respectively. The evaluation of the second [5] Fiaschi D, Manfrida G, Maraschiello F. Design and performance prediction of
radial ORC turboexpanders. Appl Energy 2015;138:51732.
law efficiency for the ORC is shown in Fig. 22b at design operating
[6] Hu D, Zheng Y, Wu Y, Li S, Dai Y. Off-design performance comparison of an
conditions (Table 3) and for both configurations. The maximum organic Rankine cycle under different control strategies. Appl Energy
second law efficiency was 75.54% and 57.22% for the two and 2015;156:26879.
single-stage configurations respectively, with n-pentane as a work- [7] Hu D, Li S, Zheng Y, Wang J, Dai Y. Preliminary design and off-design
performance analysis of an Organic Rankine Cycle for geothermal sources.
ing fluid. Energy Convers Manage 2015;96:17587.
These results from the two-stage axial turbine are better than [8] Rahbar K, Mahmoud S, Al-Dadah RK, Moazami N. Parametric analysis and
other studies in the literature, as in Refs. [6,7,27,29,30] with a max- optimization of a small-scale radial turbine for Organic Rankine Cycle. Energy
2015;83:696711.
imum ORC thermal efficiency of about 9.8% based on a two-stage [9] Rahbar K, Mahmoud S, Al-Dadah RK, Moazami N. Modelling and optimization
radial-inflow turbine with an expansion ratio of 11.6 and mass of organic Rankine cycle based on a small-scale radial inflow turbine. Energy
flow rate of 1.52 kg/s, as reported in [30]. They highlight the poten- Convers Manage 2015;91:18698.
[10] White M, Sayma AI. System and component modelling and optimisation for an
tial of this integrated approach for further accurate prediction of efficient 10 kWe low-temperature organic Rankine cycle utilising a radial
ORC performance using a small-scale two-stage axial turbine dri- inflow expander. Proc Inst Mech Eng, Part A: J Power Energy
ven by a low-temperature heat source. 2015;229:795809.
[11] Li Y, Ren XD. Investigation of the organic Rankine cycle (ORC) system and the
radial-inflow turbine design. Appl Therm Eng 2016;96:54754.
8. Conclusions [12] Song J, Gu CW, Ren X. Influence of the radial-inflow turbine efficiency
prediction on the design and analysis of the Organic Rankine Cycle (ORC)
system. Energy Convers Manage 2016;123:30816.
The organic Rankine cycle is one of the most promising tech- [13] Harinck J, Pasquale D, Pecnik R, van Buijtenen J, Colonna P. Performance
nologies that can be utilized for efficient power generation applica- improvement of a radial organic Rankine cycle turbine by means of automated
tions from low-grade heat sources such as solar and geothermal computational fluid dynamic design. Proc Inst Mech Eng, Part A: J Power
Energy 2013;227:63745.
energy. The organic Rankine cycle systems efficiency is generally [14] Sauret E, Gu Y. Three-dimensional off-design numerical analysis of an organic
low due to the low temperature difference between the heat Rankine cycle radial-inflow turbine. Appl Energy 2014;135:20211.
source and heat sink. Therefore, high organic Rankine cycle system [15] Fiaschi D, Innocenti G, Manfrida G, Maraschiello F. Design of micro radial
turboexpanders for ORC power cycles: from 0D to 3D. Appl Therm Eng
efficiency requires an efficient turbine because its performance has 2016;99:40210.
a substantial influence on the systems thermal efficiency. [16] Russell H, Rowlands A, Ventura C, Jahn I. Design and testing process for a 7 kW
This work developed a novel efficient small-scale two-stage radial inflow refrigerant turbine at the University of Queensland. Proceedings
of ASME Turbo Expo 2016: GT2016-58111, Seoul, South Korea.
axial turbine for low temperature and small scale organic Rankine
[17] Nithesh KG, Chatterjee D. Numerical prediction of the performance of radial
cycle systems. In this work, an integrated methodology of the pre- inflow turbine designed for ocean thermal energy conversion system. Appl
liminary mean-line design, three-dimensional computational fluid Energy 2016;167:16.
[18] Moroz L, Kuo CR, Guriev O, Li YC, Frolov B. Axial turbine flow path design for an
dynamics analysis and organic Rankine cycle modelling using
organic Rankine cycle using R-245fa. ASME Turbo Expo 2013: GT2013-94078,
REFPROP and real gas formulations of various working fluids was Texas, USA.
used. Results showed that the two-stage axial turbine configura- [19] Da Lio L, Manente G, Lazzaretto A. New efficiency charts for the optimum
tion exhibited substantially higher turbine performance, with design of axial flow turbines for organic Rankine cycles. Energy
2014;77:44759.
overall isentropic efficiency of 83.94%, power output of [20] Da Lio L, Manente G, Lazzaretto A. Predicting the optimum design of single
16.037 kW and ORC thermal efficiency of 14.19%, compared to stage axial expanders in ORC systems: is there a single efficiency map for
78.30%, 11.06 kW and 10.5% respectively from the single-stage different working fluids? Appl Energy 2016;167:4458.
[21] Cho SY, Cho CH, Ahn KY, Lee YD. A study of the optimal operating conditions in
configuration with n-pentane as a working fluid. The results also the organic Rankine cycle using a turbo-expander for fluctuations of the
showed that the developed small-scale two-stage turbine can be available thermal energy. Energy 2014;64:90011.
used for low-temperature (<100 C) heat source with mass flow [22] Al Jubori A, Al-Dadah RK, Mahmoud S, Khalil KM, Bahr Ennil AS. Development
of efficient small scale axial turbine for solar driven organic Rankine cycle.
rates ranging between 0.2 and 0.5 kg/s for three organic fluids (n- Proceedings of ASME Turbo Expo 2016, Seoul, South Korea, paper no. GT2016-
pentane, R141b and R45fa). These results highlight the potential 57845.
of using two-stage axial turbine to enhance the performance of a [23] Al Jubori A, Daabo A, Al-Dadah RK, Mahmoud S, Ennil AB. Development of
micro-scale axial and radial turbines for low-temperature heat source driven
small scale ORC system using low grade heat source. Finally,
organic Rankine cycle. Energy Convers Manage 2016;130:14155.
three-dimensional computational fluid dynamics optimization [24] Clemente S, Micheli D, Reini M, Taccani R. Bottoming organic Rankine cycle for
integrated with structural analysis will be developed and applied a small scale gas turbine: a comparison of different solutions. Appl Energy
2013;106:35564.
on these turbines configurations in order to enhance the turbine
[25] Chang JC, Hung TC, He YL, Zhang W. Experimental study on low-temperature
and organic Rankine cycle systems performance. organic Rankine cycle utilizing scroll type expander. Appl Energy
2015;155:1509.
Acknowledgement [26] Ziviani D, Gusev S, Lecompte S, Groll EA, Braun JE, Horton WT, et al.
Characterizing the performance of a single-screw expander in a small-scale
organic Rankine cycle for waste heat recovery. Appl Energy 2016;181:15570.
The main author (Ayad M. Al Jubori) gratefully acknowledges [27] Pei G, Li J, Li Y, Wang D, Ji J. Construction and dynamic test of a small-scale
the Iraqi Ministry of Higher Education and Scientific Research/ organic Rankine cycle. Energy 2011;36:321523.
[28] Ssebabi B, Dobson RT, Sebitosi AB. Characterising a turbine for application in
University of Technology, Baghdad, Iraq for funding the PhD schol- an organic Rankine cycle. Energy 2015;93:161732.
arship at the University of Birmingham, UK which facilitates con- [29] Kang SH. Design and experimental study of ORC (organic Rankine cycle) and
tinuation of research on the modelling and 3D analysis of small- radial turbine using R245fa working fluid. Energy 2012;41:51424.
[30] Kang SH. Design and preliminary tests of ORC (organic Rankine cycle) with
scale axial turbine. two-stage radial turbine. Energy 2016;96:14254.
[31] Cho SY, Cho CH, Choi SK. Experiment and cycle analysis on a partially admitted
References axial-type turbine used in the organic Rankine cycle. Energy 2015;90:64351.
[32] Pu W, Yue C, Han D, He W, Liu X, Zhang Q, et al. Experimental study on Organic
Rankine cycle for low grade thermal energy recovery. Appl Therm Eng
[1] Sauret E, Rowlands AS. Candidate radial-inflow turbines and high-density
2016;94:2217.
working fluids for geothermal power systems. Energy 2011;36:44607.
[33] Sung T, Yun E, Kim HD, Yoon SY, Choi BS, Kim K, et al. Performance
[2] Fiaschi D, Manfrida G, Maraschiello F. Thermo-fluid dynamics preliminary
characteristics of a 200-kW organic Rankine cycle system in a steel processing
design of turbo-expanders for ORC cycles. Appl Energy 2012;97:6018.
plant. Appl Energy 2016;183:62335.
[3] Pan L, Wang H. Improved analysis of Organic Rankine Cycle based on radial
[34] Balje OE. Turbomachines: a guide to design selection and theory. New
flow turbine. Appl Therm Eng 2013;61:60615.
York: John Wiley & Sons; 1981.
[4] Costall AW, Hernandez AG, Newton PJ, Martinez-Botas RF. Design
[35] Wilson DG, Korakianitis T. The design of high-efficiency turbomachinery and
methodology for radial turbo expanders in mobile organic Rankine cycle
gas turbines. New Jersey: Prentice Hall; 2014.
applications. Appl Energy 2015;157:72943.
A.M. Al Jubori et al. / Energy Conversion and Management 144 (2017) 1833 33

[36] Japikse D, Baines N. Introduction to turbomachinery. Concepts ETI, Inc. and [39] Klein SA. Engineering equation solver, in F-chart Software. WI: Middleton;
Oxford University Press; 1994. 2013.
[37] Moustapha H, Zelesky MF, Baines NC, Japikse D. Axial and radial [40] Kofskey MG, Nusbaum WJ. Aerodynamic evaluation of two-stage axial flow
turbines. White River Junction: Concepts NREC; 2003. turbine designed for brayton-cycle space power system. NASA TN D-4382;
[38] Dixon SL, Hall C. Fluid mechanics and thermodynamics of 1968.
turbomachinery. Oxford, UK: Butterworth-Heinemann; 2013.

You might also like