Download as pdf or txt
Download as pdf or txt
You are on page 1of 32

Accepted Manuscript

Performance enhancement of a small-scale organic Rankine cycle radial-inflow


turbine through multi-objective optimization algorithm

Ayad M. Al Jubori, Raya Al-Dadah, Saad Mahmoud

PII: S0360-5442(17)30776-4

DOI: 10.1016/j.energy.2017.05.022

Reference: EGY 10827

To appear in: Energy

Received Date: 03 January 2017

Revised Date: 04 April 2017

Accepted Date: 04 May 2017

Please cite this article as: Ayad M. Al Jubori, Raya Al-Dadah, Saad Mahmoud, Performance
enhancement of a small-scale organic Rankine cycle radial-inflow turbine through multi-objective
optimization algorithm, Energy (2017), doi: 10.1016/j.energy.2017.05.022

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form.
Please note that during the production process errors may be discovered which could affect the
content, and all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

Highlights

 Small-scale radial-inflow turbine was designed and optimized.

 ORC system modeling and optimization of radial-inflow turbine were integrated.

 The 3D optimization based on Multi-objective genetic algorithm was conducted.

 Higher turbine and thermal system efficiencies achieved with optimized turbine.
ACCEPTED MANUSCRIPT

1 Performance enhancement of a small-scale organic Rankine cycle radial-inflow turbine


2 through multi-objective optimization algorithm
3

4 Ayad M. Al Jubori a,b *, Raya Al-Dadah a, Saad Mahmoud a

5 a The University of Birmingham, School of Engineering, Edgbaston, Birmingham, B15 2TT, UK


6 b University of Technology, Baghdad, Iraq

7 *Email: ama232@bham.ac.uk, ayadms@gmail.com

8
9 Abstract
10 An effective methodology that encompasses a mean-line design, three-dimensional CFD analysis and optimization

11 and ORC system modelling of the small-scale ORC radial-inflow turbine is presented. Three-dimensional CFD

12 analysis and a multi-objective optimization algorithm were achieved using ANSYS®17 CFX and Design Exploration

13 based on 3D RANS with a k-omega SST turbulence model. The 3D optimization technique combines a design of the

14 experiment, a response surface method and multi-objective method. The optimization of the blade geometry was

15 performed using 20 design points for both nozzle and rotor blades, based on the B-splines’ technique to represent the

16 blade angles and thickness distribution. The number of blades and rotor tip clearance were included as design

17 parameters. The isentropic efficiency and power output were introduced as an optimization objective with two

18 organic working fluids, namely isopentane and R245fa. The results of the optimized geometry with R245fa showed

19 that the turbine’s and cycle’s thermal efficiencies were higher by 13.95% and 17.38% respectively, compared with a

20 base-line design with a maximum power output of 5.415 kW. Such methodology is proved to be effective as it

21 allows the enhancing of the turbine’s and the ORC’s system performance throughout to find the optimum blade

22 shape of the turbine stage.

23 Keywords: small-scale radial-inflow turbine; preliminary mean-line design; ORC; 3D CFD optimization; multi-

24 objective genetic algorithm.

25 1. Introduction

26 One of the major challenges in the world today is the increasing energy demand. Therefore, more attention is

27 dedicated to energy saving and reduction of environmental pollution and fossil fuel consumption by exploiting

28 renewable energy sources. The Organic Rankine Cycle (ORC) provides electricity from low-temperature heat

29 sources including renewable energy (i.e. solar and geothermal) and low-grade waste heat. In this scenario, the ORC

1
ACCEPTED MANUSCRIPT

30 systems offer potential for generating electricity for wide range of applications including domestic and remote off-

31 grid communities.

32 Many researches have been carried out regarding ORC system by focusing on thermodynamic analyses,

33 optimization and the selection of a suitable working fluid for the cycle for low-temperature heat source applications,

34 including solar thermal energy as reported in [1,2,3]. While, several studies focused on thermo-economic, and ORC

35 system analysis driven by geothermal energy [4,5,6,7]. In terms of low-grade waste heat sources, ORC

36 thermodynamic analysis has been conducted in [8,9,10]. In aforementioned studies, the thermodynamic analysis

37 model of ORC’s system was performed based on the assumption of constant turbine isentropic efficiency for

38 different working fluid and various operating conditions that lead inaccurate ORC performance.

39 In an ORC system, the turbine’s efficiency has a significant influence on the ORC system’s performance. For

40 small-scale power generation using ORC, radial-inflow turbine is considered a suitable choice as reported in

41 literature. A number of studies using radial-inflow turbine based on various approaches are summarized in Table 1

42 using the mean-line design approach and CFD analysis for developing radial-inflow turbine.

43 Although there have been a number of recent attempts to develop the turbine’s performance using

44 computational fluid dynamic techniques like ANSYS CFX/Fluent, a limited amount of work has involved using

45 optimization techniques to optimize the blade geometry to improve both the turbine’s and ORC’s performance. Al

46 Jubori et al. [28] presented a new methodology that coupled 1D, 3D CFD analysis and optimization of a small-scale

47 axial turbine with ORC system modelling, based on a multi-objective genetic algorithm (MOGA) with six different

48 working fluids. Their optimization results exhibited that the maximum turbine and cycle efficiencies and power

49 output with R123 were about 88%, 10.5% and 6.3 kW respectively. Rahbar et al. [29] carried out 3D optimization of

50 the transonic rotor of a two-stage radial-inflow turbine using a genetic algorithm (GA) with R245fa as the working

51 fluid. Their optimization results showed that the maximum turbine isentropic efficiency was of 88% with a

52 maximum power output of 26.35 kW and cycle efficiency about 14.8%, with a pressure ratio of 10 and total inlet

53 temperature of 405.3 K.

54 There is limited literature concerning the design and 3D CFD analysis and optimization of small-scale radial-

55 inflow turbine for ORC’s system with power output around 5 kW for different power generation applications, such

56 as small buildings, rural areas, off-grid zones and isolated installations. Therefore, three-dimensional CFD

57 optimization using multi-objective optimization for a small-scale radial-inflow turbine stage (nozzle and rotor) is

2
ACCEPTED MANUSCRIPT

58 novel and has only received limited investigation previously. New methodology for integrating the mean-line

59 design, 3D CFD analysis, and multi-objective optimization with ORC modelling has been presented for the small-

60 scale radial-inflow turbine (RIT) stage. Furthermore, it seeks to fill the gap by investigating the turbine’s

61 performance in both design and off-design conditions for baseline and optimum design cases with two organic

62 working fluids. The mean-line design of the RIT and ORC modelling is developed using the engineering equation

63 solver (EES) software; ANSYS®17-CFX is employed to predict the 3D viscous flow and turbine performance. The

64 real gas formulation of the working fluids is applied to perform an accurate prediction of the real behaviour of the

65 working fluids in a turbine/ORC model using the REFPROP database. The CFD baseline design of the RIT is

66 optimized using the ANSYS®17-Design Exploration package for 3D optimization purposes, based on a multi-

67 objective genetic algorithm (MOGA). The optimized turbine performance (isentropic efficiency and power output)

68 for each organic working fluid is inserted into the ORC model to determine the best cycle efficiency. The inclusion

69 of constraints in the optimization technique allows for achieving the highest efficiency from optimized geometry

70 without exceeding input operating conditions.

71 2. Working fluids selection and ORC system modelling

72 The selection of the organic working fluid is an essential aspect in the ORC system modelling and

73 performance. According to their thermo-physical properties, the working fluids have a strong influence on the ORC

74 system’s efficiency, the expander’s performance, the components’ size, the system’s stability, safety and economic

75 feasibility and the environmental concerns [30]. Organic working fluids have large molecular weights and a low

76 boiling temperature and pressure and are usually heavy mixtures. Based on their slopes of saturation vapour on the

77 T-s diagram, organic fluids are categorized into dry, isentropic and wet fluids. For low-temperature heat sources, the

78 dry and isentropic working fluids with positive and zero slopes (dT/ds) are favourable because of the elimination of

79 the need for the superheating condition after the expansion process, as presented in Fig. 1a. Also, the working fluids

80 should have low flammability and corrosion features and be environmentally friendly, with zero ozone depletion

81 potential (ODP) and low global warming potential (GWP); as clarified in Table 2 which illustrates the properties of

82 the selected working fluids. Due to the selection of working fluids is a main challenge for ORC turbines designers

83 and it is based on an acceptable balance between the aforementioned criteria, environmental concerns,

84 thermodynamic performance, commercial availability and cost. Therefore, isopentane and R245fa are selected based

3
ACCEPTED MANUSCRIPT

85 on these criteria and is recommended in literature as a suitable working fluid for low-temperature heat sources

86 application [30,31].

87 The main five components of a recuperative ORC are the evaporator, turbine, recuperator, condenser and a

88 pump, as shown in Fig. 1b. In this study, a subcritical ORC cycle is investigated to avert the safety concerns of the

89 system complexity of high-pressure systems. The losses of heat and pressure through the ORC piping system are

90 neglected with assumption of steady state operating conditions. Heat added from the low-temperature heat source is

91 given by:

92 Qe = m(h1 ‒ h6) (1)Net power

93 output from the ORC cycle is given by:

94 Wnet = Wtηmechηgen ‒ Wp (2)

95 where ηmech and ηgen are mechanical efficiency and generator efficiency. ORC thermal efficiency is given by:

Wnet
96 ηth = (3)
Qe

97 The second law efficiency can be defined as the proximity of the real thermal efficiency of the cycle to the

98 Carnot cycle efficiency as:

ηth Wnet
ηsec eff = = (4)
ηCarnot
( )
TL
99 Qe 1 ‒ The input
TH

100 parameters values of ORC system modelling are presented in Table 3 in terms of heat source and heat sink

101 temperature and mass flow rate with two working fluids.

102

103 3. Radial-inflow turbine (RIT) design

104 The first and crucial step of the whole turbine design procedure is the preliminary mean-line design (PD) to

105 create the correct aerodynamic design that delivers the desired output. The PD of the RIT is based on one-

106 dimensional (1D) mean-line flow analysis. The mean-line model refers to the mid-span values of the blades’ passage

107 and only focuses on the inlet and outlet condition of each blade’s passage. Its approach allows fast prediction for

108 fluid-dynamic development and thermodynamic process inside the blade’s passage. The aim of mean-line design is

4
ACCEPTED MANUSCRIPT

109 to deliver the initial dimensions of the turbine’s and blade’s shape such as hub and tip diameters, chord length, blade

110 pitch, number of blades, thickness of leading edge and thickness of trailing edge [32].

111 The RIT stage consists of three main parts, which are the volute, nozzle and rotor, as shown in Fig 2a; and the

112 velocity triangles and basic geometry of the rotor as shown in Fig. 2b, which is implemented in accordance with the

113 methodology outlined in [33, 34]. The flow and loading coefficients (ϕ, ψ) are required and assumed in order to give

114 the best efficiency. The rotor inlet blade angle (β4) is assumed equal to 70° as recommended by [34]. The present

115 model of the RIT rotor includes the losses due to incidence, passage, trailing edge, tip clearance and windage, which

116 are detailed in Table 4. The number of rotor blades is calculated based on the flowing empirical correlation [35]:

π
117 Zrotor = (110 ‒ α4)tan (α4) (5)
30

118 Based on the literature, the rotor losses are more significant in comparison with nozzle losses. Thus, the

119 isentropic expansion is assumed through the nozzle. The nozzle geometry was carried out for matching dimensions

120 as in the following equation:

121 ∆r = 2b3cos (α4) (6)

122 where b3 is the nozzle blade width.

123 The number of blades is calculated through the following relationship and the nozzle solidity equals to 1.35 as

124 reported in Glassman [35]:

2πr1
125 ZNozzle = (7)
1.35

126 The friction loss through the nozzle is calculated as [34]:


lhyd
127 ∆hfriction = 4.fnozzle.C (8)
dhyd

128 The performance of the RIT turbine is estimated based on losses of rotor, nozzle and volute in terms of

129 enthalpy drop, which allows the calculation of turbine isentropic efficiency within a highly comprehensive iterative

130 process. The total-to-total efficiency and total-to-static efficiency are defined as the following [14]:

∆hactual
131 η= (9)
∆hactual + ∑(∆h total losses)

132
133

5
ACCEPTED MANUSCRIPT

134

135 3.1 Input/output of RIT preliminary mean-line design

136 In this paper, the PD code is developed using EES software (engineering equation solver software) [39]. The

137 EES code has the ability to deliver a wide range of RIT configurations by accomplishing comprehensive studies in

138 terms of different input parameters and working fluids, as offered in Table 5. The output of the PD methodology

139 delivers the turbine layout in terms of turbine diameters and blade dimensions and shapes and is presented

140 in Table 6. The flow chart of the PD code is presented in Fig. 3 where the PD methodology is a highly iterative

141 procedure.

142

143 4. 3D CFD analysis

144 The actual configuration of the flow inside the turbine stage is particularly complex; so it requires a high-

145 fidelity model based on an adequately complex flow scheme. Therefore, the integrated methodology between the

146 low-fidelity model (i.e. PD model) and the high-fidelity model (i.e. 3D CFD model) is essential to directly predict

147 the most relevant flow features (3D, turbulent and unsteady flow etc.). Consequently, the main turbine stage’s

148 geometric characteristics from the mean-line design (i.e. exit tip and hub radii, inlet tip diameter, blade height and

149 angles), as listed in Table 6, are exported as inputs into the blade design module to generate the 3D geometry of the

150 nozzle and rotor blades, using ANSYS®17-BladeGen software as shown in Fig. 4. The pressure/suction and

151 angle/thickness modes are employed in creating the 3D blade geometry of the nozzle and rotor respectively. The

152 second phase is the computational grid generation that is automatically constructed through the ANSYS®17-

153 TurboGrid. The structured grid has hexagonal elements based on the O-H grid. Automatic topology and meshing

154 (ATM optimized) is applied to allow the choice of an appropriate topology for the blade passage based on the blade

155 angle, and the type of the leading and the trailing edge. The grid independency study is performed for the base-line

156 design for each working fluid to ensure that the CFD result is independent of the number of mesh nodes for each

157 working fluid. Two key decision parameters (i.e. turbine isentropic efficiency and dimensionless distance from the

158 wall y+) are considered in the mesh dependency study. The mesh is refined and the CFD simulation is re-run and

159 repeated until the mesh independence solution is reached, as displayed in Fig. 5. As shown, the CFD solution

160 becomes independent of the mesh elements number after 950,000 elements. The 3D RANS equations with the k-ω

161 SST turbulence model equations are solved using ANSYS®17-CFX. The k-ω SST has the advantage of using

6
ACCEPTED MANUSCRIPT

162 automatic near-wall treatment by locating the dimensionless distance (y+) for the first node after the wall to capture

163 the turbulence closure. The k-ω SST turbulence model is considered for flow separation under an adverse pressure

164 gradient, which accounts for the transport of the turbulent shear stress. Turbulence intensity at the inlet was

165 maintained at 5% as the recommended value, when no information was available about the inlet’s turbulence as

166 suggested in CFX-Solver modelling guide [41]. The k-ω transport equations carried out to find the turbulent kinetic

167 energy and the specific dissipation rate are:

168

∂𝑡
(𝜌𝑘) +

∂𝑥𝑖
(𝜌𝑘𝑢𝑖) =

( )
Γ𝑘
∂𝑘
∂𝑥𝑗 ∂𝑥𝑗
+ 𝐺𝑘 ‒ 𝑌𝑘 + 𝑆𝐾 (10)

169

∂𝑡
(𝜌𝜔) +

∂𝑥𝑖
(𝜌𝜔𝑢𝑖) =

Γ
∂𝜔
( )
∂𝑥𝑗 𝑘∂𝑥𝑗
+ 𝐺𝜔 ‒ 𝑌𝜔 + 𝑆𝜔 (11)

170 where Gk and G represent the generation of turbulent kinetic energy and its dissipation rate; Yk and Y represent the

171 fluctuating dilation in compressible turbulence; Sk and S are the source terms of the k- turbulence model.

172 The 3D CFD model of the small-scale radial-inflow turbines is considered as steady state 3D viscous, turbulent

173 and compressible and a single phase flow is applied. Also, all the walls are set to be smooth, non-slip and with

174 adiabatic conditions. The inlet boundary conditions at the nozzle inlet are set as the inlet’s total pressure and

175 temperature, with static pressure at the rotor outlet. In order to deliver a connection across the stationary and rotating

176 blades’ rows, a stage interface (mixing-plane) model is chosen at the nozzle-rotor interface and generalized grid

177 interface (GGI) is applied for stage (mixing-plane) analysis and the steady state flow. In all the CFD analyses, the

178 high resolution advection scheme was applied due to the topology and type of mesh, alongside its numerical

179 stability. The convergence was achieved; the maximum RMS was normally no higher than 10-5 (i.e. 1e-5) in all

180 simulations for the mass, momentum, energy, and turbulence model. For the most engineering applications, the

181 RMS of 1e-5 is considered as a good convergence and sufficient as recommended in CFX-Solver modelling guide

182 [41]. The real fluid properties’ database in ANSYS®17-CFX is integrated with REFPROP software, to obtain an

183 accurate thermodynamic model of the organic fluids’ properties in the 3D CFD analysis of ORC turbines.

184 The developed preliminary mean-line design (PD) of the RIT (detailed in section 3) is validated against a

185 published benchmark case, namely the Glassman case (i.e. code) as detailed in [40] and ANSYS-Vista RTD (1D

186 radial-inflow turbine) design software. The PD and CFD base-line design results in terms of total-to-total efficiency

7
ACCEPTED MANUSCRIPT

187 and power output (i.e. the global performance parameters) are in good agreement with the Glassman case and Vista

188 RTD and the deviations were within the acceptable margin for all working fluids as verified in Fig. 6.

189 Also, the 3D CFD results are validated with a real case (Jones case [42]); considering that most of the

190 range/values in Table 7 are taken from the above-mentioned reference and allow a fair comparison. Fig. 7 presents

191 the comparison results with ref. [42] which only mentions the details of the radial-inflow turbine geometry and

192 operating conditions. The maximum deviation in terms of isentropic efficiency was about 5.01%.

193

194 5. Three-dimensional optimization methodology

195 The current optimization technique combines the design of experiment (DoE) technique, response surface

196 method (RSM) and a multi-objective genetic algorithm (MOGA) through the ANSYS-Design of Exploration (DE)

197 that is integrated with 3D CFD analysis. In this methodology, the DoE is used to fill the design space throughout to

198 specify the location of sample design points that detects their space distribution for the blade geometry as design

199 parameters in efficient way and then feed the response surface method (RSM). The design of experiment (DoE)

200 technique is used to construct the database of blade configurations that were initially tested by CFD. Also, DoE

201 allows the delivery of the required information about a design point with a minimum number of sampling points.

202 The DoE can be defined as a provider of database information to a meta-model (surrogate model) that indicates the

203 dependency of turbine performance on the variation in blade shapes. To achieve the equal distribution of the design

204 parameters, the optimal-space-filling design (OSFD) technique in the DoE method was selected throughout the 3D

205 CFD optimization. The OSFD is the Latin hypercube sampling that is used with post-processing throughout the

206 design space within the DoE, aiming to achieve maximum vision with a small number of the sample design points

207 [43]. The surrogate model (i.e. RSM) strongly depends on the sample points which are obtained from the DoE.

208 Therefore the DoE was integrated with the 3D CFD optimization algorithm to deliver an efficient design space of

209 the design points (i.e. maximum discernment and minimum number of sample points) [44]. The details of the 3D

210 CFD optimization procedure are shown in Fig. 8.

211 The response surface method (RSM) was constructed as a surrogate model for each objective function based on

212 the CFD solutions achieved at specified control design points. The RSM is a methodology of fitting a polynomial

213 function for discrete responses obtained from DoE. It reveals the association between response functions and design

214 parameters. The second-order polynomial response can be defined as the following [45]:

8
ACCEPTED MANUSCRIPT

N N N
215 f(x) = β0 + ∑j = 1βjxj + ∑j = 1βjjx2j + ∑∑i ≠ jβijxixj (12)

216 Where, f(x) is a target function of optimization; x represents a group of design parameters and β is a coefficient of

217 regression.

218 The multi-objective optimization algorithm is expressed as:

219 Maximize/Minimize f (x)=((f1 (x), f2 (x), f3 (x)…. FM (x)) (M functions to be optimized)

220 Subject to g(x) ≤ 0 (inequality constraints)

221 or j(x) = 0 (equality constraints)

222 The blade geometry was parameterized as a unique B-spline means composed of 20 control points, achieving

223 through the DoE technique a high degree of geometric flexibility and robustness. The design points are the variables

224 used to define the blade geometry; while the turbine performance (i.e. turbine isentropic efficiency and power

225 output) is defined as the objective function. The 3D CFD optimization is aimed to enhance the turbine performance

226 by modifying the blade geometry, as shown in Fig. 9, to reduce the passage losses (i.e. entropy generation) and

227 minimize the non-uniformity of flow (i.e. secondary flow). The 3D optimization can be achieved through the

228 following steps: select the design points (blade geometry); define the objective function that determines the radial-

229 inflow turbine’s performance, composed with any constraints; and explore the algorithm to determine the optimum

230 objective function (i.e. turbine performance) corresponding to the optimum design points of the blade geometry.

231 The parameterization of blade geometry is a crucial aspect of an efficient optimization technique; where it is

232 necessary to create an extensive range of accurate geometries within a minimum group of design points

233 simultaneously, with a careful selection of the design points of the blade geometry and their ranges. In this 3D CFD

234 optimization, the parameterization of the blade geometry was conducted for the nozzle and rotor (i.e. turbine stage)

235 of a radial-inflow turbine.

236 As shown in Fig. 10, B-splines in fourth and third order curves are used to represent the blade angle and

237 thickness distribution in meridional direction (M) where M is non-dimensional distance from leading edge to trailing

238 edge. In the 3D optimization algorithm, the control points’ (CP) coordinates were employed and their possible

239 changes in the axial (i.e. meridional) and radial directions, as presented by the arrows in Fig. 10. The blade camber

240 line from leading edge (LE) to trailing edge (TE) is defined by blade angle distribution and thickness distribution

241 respectively. Starting from leading edge to trailing edge (LE to TE), the blade angle (i.e. β) is parameterized using a

242 B-spline curve technique with seven control point (CPa1 to CPa7) as illustrated in Fig10 a,b. Also, the blade thickness

9
ACCEPTED MANUSCRIPT

243 distribution is parameterized by seven control points (CPt1 to CPt7); with two points were fixed at leading edge and

244 trailing edge. As can be seen in Fig. 10, the arrows represent the movement direction for control points where the

245 control points of blade angle can move only in radial direction while the control points of thickness distribution can

246 move in both directions (i.e. radial direction and meridional direction from LE to TE).

247 Also, the number of the blades for both the nozzle and rotor was defined as a design parameter in all

248 optimization processes, to allow the optimizer to determine the optimum number of blades. The turbine performance

249 (i.e. turbine isentropic efficiency and power output) was defined as the objective function to be maximized in the

250 MOGA optimization technique. While the mass flow rate and rotational speed were specified as two constraints.

251 The 3D CFD optimizations were conducted using an Intel® CPU core i7 - 4820@ 3.70GHz with 48 GB RAM in

252 parallel run with 4 CPU cores. Typically, each optimization run takes between 72-96 hours.

253

254 5.1 3D optimization results

255 To obtain the optimum design for the blade geometry of a small-scale RIT, the 3D optimization is carried out

256 at nominal operating conditions for both working fluids (Table 5). The blade geometry parameterization is

257 conducted for both the nozzle and rotor with a number of blades. In the optimization process, the mass flow rate and

258 rotational speed were defined as two constraints. Therefore, the values of the optimum design points delivered from

259 3D CFD optimization for the blade geometry of the nozzle and rotor for each working fluid are presented in Table 8.

260 While, the original shape from the base-line design of the nozzle and rotor blades’ shapes compared to the optimum

261 design for both working fluids is illustrated in Fig. 11.

262 The spline curve in fourth and third order form was used to define the blade angle and thickness distribution as

263 design points/parameters, with the assumption of the uniform distribution of the blade thickness. In the DoE method

264 and based on the number of the design points of the blade geometry and their range, they were divided into 250

265 sample points throughout the optimal-space-filling design (OSFD) technique and varied continuously over specified

266 ranges. The number of design points was delivered from the effective distribution of the range of design points to

267 achieve the required number of samples of the design points throughout OSFD. As can be seen from Fig. 12, the

268 maximum isentropic efficiency and power output were 87.40% and 5.415 kW respectively for R245fa compared

269 with 84.35% and 4.832 kW for isopentane.

10
ACCEPTED MANUSCRIPT

270 The 3D CFD analyses and optimization exhibited further improvement in the qualitative performance of the

271 turbine stage (i.e. blade loading (pressure distribution), velocity vectors, velocity stream-lines and entropy

272 generation contours) as shown in Figs. 13-16. The pressure distribution i.e. blade loading throughout the rotor

273 passage in both base-line and optimum design cases is presented in Fig. 13 for both working fluids; it shows the

274 highest the values of the pressure on the pressure side. While the lowest values of pressure are positioned on the

275 suction side, due to the highest values of flow velocity at the throat area of the blade passage. The isentropic

276 enthalpy drop (work) is provided by the area circumscribed by such pressure distribution curves. Where, the

277 enclosed area is indicative of the net torque producing aerodynamic force by the rotor turbine shaft.

278 As can be seen from velocity vectors in Fig. 14 a-d for both working fluids, the base-line design suffers from

279 flow reversal compared with the optimized blade geometry. For both working fluids, the optimized blade shape has

280 a superior and smooth flow with no flow reversal, as shown by the velocity vectors and velocity stream-lines in the

281 blade passage, depicted in Figs.14 and 15 respectively.

282 It is evident in Figs. 16, at the base-line design for both working fluids; the flow separation and secondary

283 flows resulted in a considerable entropy generation. Where it is propagated downstream through the turbine stage as

284 shown in Fig. 16 in the base-line design passage, the entropy generation occupied the majority that led to substantial

285 losses and hence reduced the turbine isentropic efficiency. In the same figure, comparing the optimum design with

286 the base-line design shows that the entropy generation in the optimized geometry has substantially reduced

287 throughout the optimization process. The enhancements in the flow aerodynamics through the blade passage clarify

288 the improvement in the turbine’s performance (i.e. efficiency and power).

289 The maximum enhancement in the optimized turbine performance (isentropic total-to-total efficiency and

290 power output) is shown in Figs. 17a and b with maximum improvement in terms of efficiency of 13.95% and power

291 output of about 14% with R245fa as the working fluid; while the maximum enhancement with isopentane was of

292 10.55% and 11.98% in terms of efficiency and power output respectively. This improvement in the turbine

293 performance reveals the effectiveness of the 3D CFD optimization technique from modifying the turbine blade’s

294 geometry.

295 It is evident from Figs. 18 and 19 that the turbine performance in terms of efficiency and power output

296 improved in off-design conditions as well as with in-design conditions for a wide range of pressure ratios and

297 rotational speeds for both working fluids (R245fa and isopentane). The design point was presented in Table 5 with a

11
ACCEPTED MANUSCRIPT

298 pressure ratio of 3.0 and rotational speed of 27500 rpm for isopentane and 35000 rpm for R245fa. When the total

299 pressure ratio is increased, the enthalpy drop increases, leading to a large power output. While the turbine efficiency

300 drops with increasing the rotational speed for both working fluids based on the definition of the loading coefficient,

301 it is explained the variation in enthalpy drop (i.e.specific work) througout the radial turbine stage. These figures

302 reflect the highlighted advantages of using 3D CFD optimization techniques in off-design conditions as well.

303

304 6. ORC analysis results

305 Using the optimized turbine performance (i.e. isentropic efficiency and power output) at the design point

306 (Table 3) and setting these as inputs to the ORC model (section 2) resulted in the system’s thermal efficiency of

307 11.27%, compared with 9.56% at base-line design for R245fa as the working fluid, as shown in Fig. 20a. Also, Fig.

308 20a shows that the optimized ORC system’s efficiency with isopentane as the working fluid was 9.69% compared

309 with 8.07% at base-line design. Such results demonstrate the effectiveness of 3D CFD optimization to further

310 improve the ORC’s performance. The assessment of the ORC system’s second law efficiency with both working

311 fluids at base-line and optimum design cases is shown in Fig. 20b. The maximum second law efficency is 71.40%

312 for R245fa, compared with about 64.00% for isopentane at optimum design. As can be seen in Figs. 21 and 22,

313 clearly the optimization technique was very effective as it improved the cycle’s thermal efficiency in both design

314 and off-design conditions for various pressure ratios and total inlet temperatures for both working fluids.

315

316 7. Conclusions
317 In this paper, 3D CFD optimization of the blade’s geometry of a small-scale radial-inflow turbine stage (nozzle

318 and rotor) for a low power ORC system, driven by a low-temperature heat source has been conducted to enhance the

319 turbine’s and the ORC system’s performance. R245fa and isopentane were selected as working fluids with a

320 temperature heat source of (≈ 90 °C). In multi-objective optimization algorithm, the turbine performance (isentropic

321 efficiency and power output) was selected as an objective function subject to maximization. Two constraints (mass

322 flow rate and rotational speed) were defined to keep the operating conditions fixed during the optimization process;

323 this allows the achievement of the highest performance.

324 It was revealed from the 3D CFD MOGA optimization results that the flow aerodynamics was improved

325 significantly compared with the base-line design for both working fluids. Also, the results exhibited that R245fa has

12
ACCEPTED MANUSCRIPT

326 a maximum turbine performance with isentropic efficiency of 87.40% and power output of 5.415 kW with a thermal

327 ORC system efficiency of 11.27%. This effective integrated methodology inclusion (mean-line design, CFD

328 analysis and MOGA optimization) proved to give a remarkable improvement in the performance of the small-scale

329 ORC radial-inflow turbine stage based on the optimization of the blade’s geometry. The 3D CFD optimization

330 showed that a more efficient turbine performance can be achieved by optimizing the blade’s geometry in terms of

331 blade angles, thickness distribution of the blade and the number of blades for both the nozzle and rotor, throughout

332 assessing their quantitative performance (isentropic efficiency and power output) and qualitative aerodynamic

333 performance (velocity vectors, velocity stream-lines, pressure distribution and entropy generation contours).

334 Furthermore, such optimization results revealed the potential and effectiveness in design and off-design CFD

335 analysis for a wide range of rotational speeds and pressure ratios for both working fluids. These results highlight the

336 potential and advantages of using the 3D MOGA optimization technique to achieve high turbine efficiency leading

337 to high ORC system thermal efficiency.

338 Acknowledgement
339 The main author (Ayad M. Al Jubori) gratefully acknowledges the Iraqi ministry of higher education and scientific
340 research for funding PhD scholarship at the University of Birmingham, UK which facilitates continuation of
341 research on the modelling and 3D optimization of small-scale radial-inflow turbine.
342
343 References
344
345 [1] Wang M, Wang J, Zhao Y, Zhao P, Dai Y. Thermodynamic analysis and optimization of a solar-driven
346 regenerative organic Rankine cycle (ORC) based on flat-plate solar collectors. Applied Thermal Engineering
347 2013;50(1):816-25.
348 [2] Hajabdollahi H, Ganjehkaviri A, Jaafar MN. Thermo-economic optimization of RSORC (regenerative solar
349 organic Rankine cycle) considering hourly analysis. Energy 2015;87:369-80.
350 [3] Helvaci HU, Khan ZA. Thermodynamic modelling and analysis of a solar organic Rankine cycle employing
351 thermofluids. Energy Conversion and Management 2017;138:493-510.
352 [4] Wang X, Liu X, Zhang C. Parametric optimization and range analysis of Organic Rankine Cycle for binary-cycle
353 geothermal plant. Energy Conversion and Management 2014;80:256-65.
354 [5] Liu Q, Shen A, Duan Y. Parametric optimization and performance analyses of geothermal organic Rankine
355 cycles using R600a/R601a mixtures as working fluids. Applied Energy 2015;148:410-20.
356 [6] Kazemi N, Samadi F. Thermodynamic, economic and thermo-economic optimization of a new proposed organic
357 Rankine cycle for energy production from geothermal resources. Energy Conversion and Management
358 [7] Sadeghi M, Nemati A, Yari M. Thermodynamic analysis and multi-objective optimization of various ORC
359 (organic Rankine cycle) configurations using zeotropic mixtures. Energy 2016; 109:791-802.
360 [8] Maraver D, Royo J, Lemort V, Quoilin S. Systematic optimization of subcritical and transcritical organic
361 Rankine cycles (ORCs) constrained by technical parameters in multiple applications. Applied energy 2014;117:11-
362 29.
363 [9] Feng Y, Hung T, Zhang Y, Li B, Yang J, Shi Y. Performance comparison of low-grade ORCs (organic Rankine
364 cycles) using R245fa, pentane and their mixtures based on the thermoeconomic multi-objective optimization and
365 decision makings. Energy 2015;93:2018-29.
366 [10] Ge Z, Wang H, Wang HT, Wang JJ, Li M, Wu FZ, Zhang SY. Main parameters optimization of regenerative
367 organic Rankine cycle driven by low-temperature flue gas waste heat. Energy 2015;93:1886-95.

13
ACCEPTED MANUSCRIPT

368 [11] Sauret E, Rowlands AS. Candidate radial-inflow turbines and high-density working fluids for geothermal
369 power systems. Energy 2011;36(7):4460-7.
370 [12] Hu D, Li S, Zheng Y, Wang J, Dai Y. Preliminary design and off-design performance analysis of an Organic
371 Rankine Cycle for geothermal sources. Energy Conversion and Management 2015; 96:175-87.
372 [13] Rahbar K, Mahmoud S, Al-Dadah RK, Moazami N. Parametric analysis and optimization of a small-scale
373 radial turbine for Organic Rankine Cycle. Energy 2015; 83:696-711.
374 [14] Fiaschi D, Manfrida G, Maraschiello F. Design and performance prediction of radial ORC turbo expanders.
375 Applied Energy 2015;138:517-32
376 [15] Song J, Gu CW, Ren X. Influence of the radial-inflow turbine efficiency prediction on the design and analysis
377 of the Organic Rankine Cycle (ORC) system. Energy Conversion and Management 2016;123:308-16.
378 [16] Pan L, Wang H. Improved analysis of Organic Rankine Cycle based on radial flow turbine. Applied thermal
379 engineering 2013;61:606-15.
380 [17] White M, Sayma AI. System and component modelling and optimisation for an efficient 10 kWe low-
381 temperature organic Rankine cycle utilising a radial inflow expander. Proceedings of the Institution of Mechanical
382 Engineers, Part A: J. Power and Energy 2015;229:795-809.
383 [18] Cho SY, Cho CH, Ahn KY, Lee YD. A study of the optimal operating conditions in the organic Rankine cycle
384 using a turbo-expander for fluctuations of the available thermal energy. Energy 2014; 64: 900-11.
385 [19] Al Jubori A, Daabo A, Al-Dadah RK, Mahmoud S, Ennil AB. Development of micro-scale axial and radial
386 turbines for low-temperature heat source driven organic Rankine cycle. Energy Conversion and Management.
387 2016;130:141-55.
388 [20] Nithesh KG, Chatterjee D. Numerical prediction of the performance of radial inflow turbine designed for ocean
389 thermal energy conversion system. Applied Energy 2016; 167:1-6.
390 [21] Russell H, Rowlands A, De Miranda Ventura C, Jahn I. Design and testing process fora 7kw radial inflow
391 refrigerant turbine at the University of Queensland. Proceedings of ASME Turbo Expo 2016 Seoul, South Korea:
392 paper no. GT2016-58111, pp. V008T23A036.
393 [22] Fiaschi D, Innocenti G, Manfrida G, Maraschiello F. Design of micro radial turboexpanders for ORC power
394 cycles: From 0D to 3D. Applied Thermal Engineering 2016;99:402-10.
395 [23] Li Y, Ren XD. Investigation of the organic Rankine cycle (ORC) system and the radial-inflow turbine design.
396 Applied Thermal Engineering 2016; 96:547-54.
397 [24] Zheng Y, Hu D, Cao Y, Dai Y. Preliminary design and off-design performance analysis of an Organic Rankine
398 Cycle radial-inflow turbine based on mathematic method and CFD method. Applied Thermal Engineering 2017;
399 112:25-37.
400 [25] Pei G, Li J, Li Y, Wang D, Ji J. Construction and dynamic test of a small-scale organic rankine cycle. Energy
401 2011;36:3215-23.
402 [26] Kang SH. Design and experimental study of ORC (organic Rankine cycle) and radial turbine using R245fa
403 working fluid. Energy 2012;41(1):514-24.
404 [27] Sung T, Yun E, Kim HD, Yoon SY, Choi BS, Kim K, Kim J, Jung YB, Kim KC. Performance characteristics
405 of a 200-kW organic Rankine cycle system in a steel processing plant. Applied Energy 2016; 183: pp. 623-35.
406 [28] Al Jubori A, Al-Dadah RK, Mahmoud S, Ennil AB, Rahbar K. Three dimensional optimization of small-scale
407 axial turbine for low temperature heat source driven organic Rankine cycle. Energy Conversion and Management
408 2017;133:411-426.
409 [29] Rahbar K, Mahmoud S, Al-Dadah RK, Moazami N. One-dimensional and three-dimensional numerical
410 optimization and comparison of single-stage supersonic and dual-stage transonic radial inflow turbines for the ORC.
411 Proceedings of the ASME Power and Energy 2016: North Carolina, USA, pp. V001T08A017.
412 [30] Bao J, Zhao L. A review of working fluid and expander selections for organic Rankine cycle. Renewable and
413 Sustainable Energy Reviews 2013;24:325-42.
414 [31] Tchanche BF, Lambrinos G, Frangoudakis A, Papadakis G. Low-grade heat conversion into power using
415 organic Rankine cycles–A review of various applications. Renewable and Sustainable Energy Reviews
416 2011;15(8):3963-79.
417 [32] Wilson DG, Korakianitis T. The design of high-efficiency turbomachinery and gas turbines. 2nd ed. MIT press;
418 1998.
419 [33] Moustaph H, Zelesky MF, Baines NC, Japikse D. Axial and radial turbines. 1st ed. White River Junction:
420 Concepts NREC; 2003.
421 [34] Whitfield, A. and N.C. Baines, Design of radial turbomachines. 1st ed. New York: Longman; 1990.
422 [35] Glassman, AJ. Computer program for design analysis of radial-inflow turbines. NASA TN 8164; 1976.

14
ACCEPTED MANUSCRIPT

423 [36] Ventura CA, Jacobs PA, Rowlands AS, Petrie-Repar P, Sauret E. Preliminary design and performance
424 estimation of radial inflow turbines: An automated approach. Journal of Fluids Engineering 2012; 134: 0311021-13.
425 [37] Suhrmann JF, Peitsch D, Gugau M, Heuer T, Tomm U. Validation and development of loss models for small
426 size radial turbines. Proceedings of ASME Turbo Expo 2010: power for land, sea and Air GT2010, Glasgow, UK,
427 Paper No GT2010-22666.
428 [38] Japikse D, Baines N. Introduction to turbomachinery. 1994.
429 [39] Klein, SA., Engineering equation solver, in F-chart Software. 2013, Middleton, WI.
430 [40] Glassman AJ. Enhanced Analysis and User’s Manual for Radial-Inflow Turbine Conceptual Design Code RTD.
431 NASA, Contractor Report No. 195454, 1995.
432 [41] ANSYS®17 CFX-Solver Modelling Guide.
433 [42] Jones AC. Design and Test of a Small, High Pressure Ratio Radial Turbine. Trans. ASME J. Turbomach 1996;
434 118(2): pp. 362–370.
435 [43] Montgomery, DC. Design and Analysis of Experiments. 2006: John Wiley & Sons.
436 [44] Thévenin D, Janiga G. Optimization and computational fluid dynamics. Berlin, Heidelberg: Germany:
437 Springer-Verlag; 2008.
438 [45] Kim JH, Choi JH, Husain A, Kim KY. Performance enhancement of axial fan blade through multi-objective
439 optimization techniques. J. Mechanical Science and Technology 2010; 24(10): pp. 2059-66.
440
441
442
Nomenclature

b3, b4, b5 blade width (m) Subscript/superscript


C absolute velocity (m/s) 1-6 station within the turbine and cycle
d diameter (m) cr critical
f friction coefficient’s (-) e evaporator
h specific enthalpy (kJ/kg) H hot
𝑙 length (m) hyd hydraulic
K losses coefficient’s (-) L low
k specific turbulence kinetic energy (m2/s2) LE leading edge
𝑚 mass flow rate (kg/s) nbp normal boiling point
M meridional length (%) r radial
p pressure (bar) t turbine
𝑄 heat (kW) TE trailing edge
r radius (m) th thermal
s entropy (kJ/kg.K) t-t total-to-total
T temperature (K) x axial
t time (s)  tangential/circumferential direction
U blade velocity (m/s) Acronyms
w relative velocity (m/s) 1D, 3D one and three-dimensional
𝑊 power (kW) CFD computational fluid dynamics
Z number of blades (-) CP control point
GWP global warming potential
Greek symbols ODP ozone depletion potential
α absolute flow angle (degree) OSFD optimal-space-filling design
β relative flow angle (degree) MOGA multi-objective genetic algorithm
η efficiency (%) ORC organic Rankine cycle
ρ density (kg/m3) PD preliminary mean-line design
 clearance (m) RANS Reynolds-averaged Navier-Stokes
 flow coefficient (-) RSM response surface method
 loading coefficient (-) RIT radial-inflow turbine
ω specific turbulence dissipation rate (m2/s3) SST shear stress transport
443

15
ACCEPTED MANUSCRIPT

444

445 Table 1 number of studies using various investigation approaches for axial and radial-inflow turbine.
Authors Investigation Expander type Heat source Working Performance (turbine
approach temperature fluids efficiency, power, cycle
(K) efficiency)
Sauret & 1D mean-line Radial-inflow 413 R134a, 77%, 338 kW
Rowlands [11] design R143a,
R236fa,
R245fa
Hu et al. [12] 1D mean-line Radial-inflow 363 R245fa 82.3%, 66.9 kW, 5.5%
design
Rahbar et al. 1D design and Radial-inflow 362.2 8 working 84%, 15 kW
[13] optimization fluids
Fiaschi et al. 1D design Radial-inflow 420 6 working 78.8%, 50kW
[14] fluids
Song et al. [15] 1D design Radial-inflow 393.15 7 working 81%, 249.2 kW, 8.7%
fluids
Pan and Wang 1D Radial-inflow 363 14 working 75.75% ,8.15 kW, 5.2%
[16] optimization fluids
and analysis
White and 1D Radial-inflow 390 15 working 80%, 7.23 kW, 7.26%
Sayma [17] optimization fluids
and analysis
Cho et al. [18] CFD analysis Radial-inflow 393 R245fa 53% , 3.8 kW, 6.25%
for nozzle
only
Al Jubori et al. 1D and 3D Axial and 360 R141b, 83.48%, 8.507 kW,
[19] CFD analysis radial-inflow R1234yf, 10.60%
with ORC turbine R245fa, n-
modelling butane and
n-pentane
Nithesh and 1D and 3D Radial-inflow 297.5 R134a 70%%, 2 kW
Chatterjee [20] CFD analysis
Russell et al. 1D design and Radial-inflow 423 R245fa 76%, 7 kW
[21] 3D CFD
simulation for
rotor only
Fiaschi et al. 1D design and Radial-inflow 395.9 R134a 68.04%, 4.504 kW
[22] 3D CFD
simulation
Li & Ren [23] 1D, 3D design Radial-inflow 393 R123 84.33%, 534 kW
and simulation
Zheng et al. 1D, 3D design Radial-inflow 360 R134a 81.6%, 643 kW
[24] and simulation
Pei et al. [25] Experimental Radial-inflow 373 R123 65%, 1.36 kW, 6.8%
Kang [26] Experimental Radial-inflow 358.4 R245fa 78.7% ,32.7 kW, 5.22%
Sung et al. [27] Experimental Radial-inflow 413 R245fa 177.4 kW and 9.6%
446

447

448

16
ACCEPTED MANUSCRIPT

449

450 Table 2 Thermodynamic property of six organic fluids.


Working Mol. Tnbp Tcr Pcr ODP GWP
Fluid weight (K) (K) (kPa) (100 yr)
(g/mol)
Isopentane 72.149 300.98 460.35 3378 0 20
R245fa 134.05 288.14 426. 16 3610 0 950
451

452 Table 3 The input parameters of the ORC model.


Parameters Unit Value
Heat source temperature K 365
Heat sink temperature K 293
Pump efficiency - 0.75
Generator efficiency - 0.96
Mechanical efficiency - 0.96
Recuperator effectiveness - 0.8
Working fluid mass flow rate kg/s 0.3
453

454 Table 4 Losses’ modelling of RIT rotor.


Type of losses Correlation Reference
2
𝑤𝜃4 [36]
Incidence loss
∆ℎ𝑖𝑛𝑐𝑖𝑑𝑒𝑛𝑐𝑒 =
2
𝐾𝑓𝜌𝑈34𝑟24 [34]
Disk friction
∆ℎ𝑑𝑖𝑠𝑘 𝑓𝑟𝑖𝑐𝑡𝑖𝑜𝑛 =
4𝑚
𝑤5𝑡𝑖𝑝 + 𝑤5ℎ𝑢𝑏
( )
[ ]
𝑤4 + [37]
Friction loss 2 𝑙ℎ𝑦𝑑
∆ℎ𝑓𝑟𝑖𝑐𝑡𝑖𝑜𝑛 = 𝑓𝑐𝑢𝑟𝑣𝑒
2 𝑑ℎ𝑦𝑑

𝐶24.𝑑4 [34]
Secondary loss
∆ℎ𝑠𝑒𝑐𝑜𝑛𝑑𝑎𝑟𝑦 =
𝑍𝑟𝑜𝑡𝑜𝑟.𝑟𝑐
Tip clearance loss 𝑈34.𝑍𝑟𝑜𝑡𝑜𝑟
∆ℎ𝑡𝑖𝑝 𝑐𝑙𝑒𝑟𝑎𝑛𝑐𝑒 =
8𝜋
(0.4.𝜀𝑥.𝐶𝑥 + 0.75.𝜀𝑟.𝐶𝑟 ‒ 0.3 𝜀𝑥.𝜀𝑟.𝐶𝑥.𝐶𝑟) [38]

𝜌.𝑈4.3 𝑟24 [33,36]


Windage loss
∆ℎ𝑤𝑖𝑛𝑑𝑎𝑔𝑒 = 𝐾𝑓
2.𝑚.𝑤25
Exit energy loss ∆ℎ𝑒𝑥𝑖𝑡 = 0.5𝐶25 [37]
455

456 Table 5 Input parameters of PD for radial-inflow turbine and their ranges/values.
Parameters Values/ Ranges
Loading coefficient () (-) 0.6-1.8
Flow coefficient () (-) 0.1-0.6
Nozzle RIT inlet to exit radius ratio (r2/r3) (-) 1.2-1.3
Rotor RIT exit hub to inlet radius ratio (r5hub/r4) (-) 0.4-0.6
Rotor exit absolute flow angle (α5) (degree) -10

17
ACCEPTED MANUSCRIPT

Blade speed ratio (-) 0.7


Rotational speed (rpm) 20000-40000
Inlet total temperature (k) 365
Inlet total pressure (bar) Corresponding saturated vapour
pressure at inlet temperature
Mass flow rate (kg/s) 0.3
Working fluids isopentane, R245fa
457 Table 6 Output of PD radial-inflow turbine mean-line design for each working fluid.
Parameter Working fluids
isopentane R245fa
α4 (degree) 72.33 66.78
β4 (degree) 48.91 9.75
βblade,4 (degree) 68.91 59.75
βtip,5 -65.47 -60.65
d1 (mm) 63.47 50.65
d2 (mm) 55.23 44.09
d3 (mm) 43.81 34.35
d4 (mm) 41.31 31.85
dtip,5 (mm) 30.27 22.48
b4 (mm) 2.84 2.05
b5 (mm) 21.72 14.51
Zrotor 15 11
Nozzle LE beta angle (degree) -20 -15
Nozzle TE beta angle (degree) 68 63
Nozzle stagger angle (degree) 33 38
Nozzle throat (mm) 3.481 3.926
Znozzle 29 33
458

459 Table 7 Details of the radial-inflow turbine geometry of a real case (Jones case [42]).
Parameters Value Parameters Value
r2 (m) 0.074 β4 (degree) -31.8
r3 (m) 0.0635 β5 (degree) -57.40
r4 (m) 0.0582 Znozzle (-) 19
r5tip (m) 0.0368 Zrotor (-) 16
r5hup (m) 0.0152 Nozzle chord(m) 0.0229
b4 (m) 0.00618 Rotor chord (m) 0.0457
b5 (m) 0.00635 Radial Clearance (m) 0.23 ×10-3
α3 (degree) 77.80 Nozzle trailing edge thickness (m) 0.51 ×10-3
α4 (degree) 76.8 Rotor trailing edge thickness (m) 0.76 ×10-3
α5 (degree) -0.03
460

461 Table 8 The optimum design parameters of the blade geometry for both working fluids.
Parameters Working fluids
isopentane R245fa
Rotor inlet absolute flow angle,α4 (degree) 76.25 69.37
Rotor outlet absolute flow angle, α5 (degree) -17.36 -11.21
Rotor inlet relative flow angle, β4 (degree) 54.75 13.71
Rotor blade angle, βblade,4 (degree) 60.58 65.64
Rotor outlet relative angle βtip,5 -55.17 -65.34
Rotor radial tip clearance (m) 0. 325×10-3 0. 345×10-3
Rotor blade chord (m) 0.0386 0.0368
Zrotor 18 14
Nozzle LE beta angle (degree) -12 -10
Nozzle TE beta angle (degree) 71 67

18
ACCEPTED MANUSCRIPT

Nozzle stagger angle (degree) 36 40


Nozzle throat (mm) 3.25 3.54
Znozzle 33 37
Nozzle blade chord (m) 0.0153 0.0147
462

Fig. 1. Working fluids T-s diagram (a), layout of recuperative ORC system (b).
463

464

Fig. 2. Stage cross section (left), velocity triangles and rotor blade profile (right) [19].
465

19
ACCEPTED MANUSCRIPT

466

467

468

469

470

471
Start
472
Boundary
473 Condition and
design parameters
(Table 5)
474

475 Guess
(Total-to-total Efficiency)
476
Calculate working
477 fluids’ properties

478
Calculate
(velocity triangle, overall
479 size of blade geometry for
both nozzle and rotor)
480

481 Calculate
new total-to-total efficiency based on
losses model (Table 4 and equation 9)
482

483 No If
|(𝒕 ‒ 𝒕)𝒏𝒆𝒘 ‒ (𝒕 ‒ 𝒕)𝒐𝒍𝒅 | < 𝟏𝟎 ‒ 𝟑

484
Yes
485
Initial dimensions and
performance of turbine
486

487 End

488 Fig. 3. Flow chart of the preliminary mean-line design.

489

490

20
ACCEPTED MANUSCRIPT

491

492

493

494

495

496
Rotor
497
Nozzle
498

499

500

501

502

503 Fig. 4. Three-dimensional radial-inflow turbine stage (left), meshes generation for nozzle and rotor blade passage (right).

504

505

506

507

508

509

510

511

512

513

514 Fig. 5. The grid independency based on total-to-static turbine efficiency with R245fa as the working fluid.

515

21
ACCEPTED MANUSCRIPT

516

517

518

519

520

521

522

523

524

525

526

527

528
Fig. 6. Comparison: the CFD and PD with Vista RTD and Glassman code in terms of turbine isentropic
529 efficiency (left) and power output (right).

530

531

532

533

534

535

536

537

538

539

540

Fig. 7. Comparison: the CFD efficiency with experimental efficiency of ref. [42].
22
ACCEPTED MANUSCRIPT

541

542

543

544

545

546

547

548

549

550

551

552

553

554

555

556

557

558

559

560

561

562

563

564

565

23

Fig. 8. Flow chart of 3D CFD optimization of radial-inflow turbine.


ACCEPTED MANUSCRIPT

566

567

568

569

570

571

572

573

574

575

576 Design points

577

578

579 Fig. 9. Definition and parameterization of the blade geometry.

580

581

582 CPa1
CPa2 CPt2 CPt3
CPa3 CPa4 CPt4
583 CPt5
CPa5 CPt1
584 CPt6
CPa6
585
CPa7 CPt7
586

587

588

589
Fig. 10. Parameterization of the distribution of blade angle (left), and blade thickness (right).
590

24
ACCEPTED MANUSCRIPT

591

592

593

594

595

596

597

598

599
Nozzle of R245fa
600 Nozzle of isopentane

601
Base-line design
602 Optimum design

603 Rotor of R245fa


Rotor of isopentane
604

605

606
Fig. 11. Nozzle and rotor blade shape at the base-line and optimum design for both working fluids.
607

608

609

610

611

612

613

614

615

25power output with design point for R245fa (left) and


Fig. 12. Variation of total-to-total efficiency and
isopentane (right).
ACCEPTED MANUSCRIPT

616

617

618

619

620

621

622

623

624

625
a) isopentane b) R245fa
626

627

628
Fig. 13. Effect of optimization on blade loading at baseline design and optimum design.
629
a) isopentane base-line design b) isopentane at optimum design
630

631

632

633

634

635

636

637 c) R245fa at base-line design d) R245fa at optimum design

638

639

640

26

Fig. 14. Velocity vectors at mid-span for base-line and optimum design with two working fluids.
ACCEPTED MANUSCRIPT

641

642

643

644

645
a) isopentane base-line design b) isopentane at optimum design
646

647

648

649

650

651

652
c) R245fa at base-line design d) R245fa at optimum design
(R245fa)
653

654

655

656

657

658
Fig. 15. Velocity streamlines at mid-span for base-line and optimum design with two working fluids.
659

660

661

662

663

664

665

27
ACCEPTED MANUSCRIPT

666

667

668

669

670
a) isopentane at base-line design
671 b) isopentane at optimum design

672

673

674

675

676

677
c) R245fa at base-line design d) R245fa at optimum design
678 (R245fa) (R245fa)

679

680

681

682

683
Fig. 16. Entropy contours at mid-span for base-line and optimum design with two working fluids.
684

685

686 a) b)

687

688

689

690

28
Fig. 17. Total-to-total isentropic efficiency (a) and power output (b) at base-line design and optimum
design for both working fluids.
ACCEPTED MANUSCRIPT

691

692

693

694

695

696 a) b)

697

698

699

700

701

702

703
Fig. 18. 3D optimization influence on the turbine performance with pressure ratio; efficiency (left) and
704 power output (right) at design and off-design conditions for both working fluids.

705

706 a) b)

707

708

709

710

711

712

713
Fig. 19. 3D optimization influence on the turbine performance with rotational speed; efficiency (left) and
714 power output (right) at design and off-design conditions for both working fluids.

715

29
ACCEPTED MANUSCRIPT

716

717

718

719

720
b)
721 a)

722

723

724

725

726
Fig. 20. The cycle’s thermal and second law efficiencies at base-line design and optimum design for both
727 working fluids.

728

729

730

731

732

733

734

735

736

737 Fig. 21. The off-design cycle’s thermal prediction Fig. 22. The off-design cycle’s thermal prediction with
with pressure ratio based on the optimized turbine inlet total temperature based on the optimized turbine
performance. performance.
738

739

740

30

You might also like