You are on page 1of 33

1 A New Approach for the Design of Diffuser-Augmented Hydro

2 Turbines Using the Blade Element Theory


3 Paulo A. S. F. Silvaa,∗, Deborah A. T. D. Rio Vazb , Vinicius Brittoa , Taygoara F. de
4 Oliveiraa,∗∗, Jerson R. P. Vazc , Antonio C. P. Brasil Juniora
a
5 University of Brası́lia - Faculty of Mechanical Engineering, Campus Darcy Ribeiro, Brası́lia, DF,
6 70910-900, Brazil
b
7 Federal University of Pará - Natural Resources Engineering Program, Av. Augusto Correa, N 1 - Belém,
8 PA, 66075-900, Brazil
c
9 Federal University of Pará - Faculty of Mechanical Engineering, Av. Augusto Correa, N 1 - Belém, PA,
10 66075-900, Brazil

11 Abstract

12 It is known that hydro turbine surrounded by a diffuser may exceed the Betz-Joukowsky limit.
13 However, it is still not clear how the diffuser efficiency affects the turbine performance. This
14 paper presents a new approach to design diffuser-augmented hydro turbines considering the
15 diffuser efficiency. Based on the blade element theory, new expressions for the axial induction
16 factor and thrust are obtained. To assess the proposed model, a comparative evaluation of
17 two different diffusers (flanged conical diffuser and flanged lens diffuser) is performed. A
18 numerical modeling investigation using computational fluid dynamics is carried out, which is
19 under an axisymmetric hypothesis based on the Reynolds Averaged Navier-Stokes technique
20 using the κ − ω shear-stress transport turbulence model. Evaluations for both turbine and
21 diffuser are performed using experimental data available in the literature. Numerical and
22 theoretical results are compared for a shrouded turbine equipped with a 83% efficiency
23 diffuser. The relative difference observed for the maximum power coefficient is about 5.3%.
24 For the hydro turbine with flanged conical diffuser, the mass flow rate is about 20% higher
25 than for a bare turbine, while for the turbine with flanged lens diffuser the increasing is only
26 2.4% more. Also, for the flanged conical diffuser the power is increased by 53%. Furthermore,
27 it is observed that the proposed blade element theory with diffuser yielded good agreement
28 with the numerical model, suggesting better performance when compared with other models
29 available in the literature.
30 Keywords: Hydrokinetic turbines, Diffuser, Blade Element Theory, Computational Fluid
31 Dynamics, Tidal turbines.


Currently at Centre for Computational Engineering sciences, Cranfield University - Cranfield MK43
0AL, United Kingdom
∗∗
Corresponding author
Email address: taygoara@unb.br (Taygoara F. de Oliveira)
Preprint submitted to Energy Conversion and Management February 19, 2018
32 1. Introduction

33 Over the past years, diffuser-augmented hydro turbines (DAHTs) have become a promis-
34 ing alternative to conventional hydropower, mainly because the diffuser technology improves
35 the turbine performance, enabling the Betz-Joukowsky limit (16/27) to be exceeded. Such
36 turbines in general do not require large flooded areas or civil structures [1, 2], contributing
37 to reduce environment impacts.
38 Several works on hydro or tidal turbines have been made in the literature [3–5], however
39 the authors are unaware of any study taking into account diffuser efficiency on the design
40 of DAHTs. Recently Silva et al. [6] developed an optimization approach based on Blade
41 Element Theory (BET) for horizontal axis hydrokinetic turbines considering the possibility
42 of cavitation, where the minimum pressure coefficient is the criterion used for identifying
43 cavitation on blades. Their model has demonstrated good behavior and indeed minimizes
44 cavitation inception. But, that work does not deal with diffusers. Gaden & Bibeau [7]
45 developed a numerical study to investigate the usage of diffusers to enhance the performance
46 and viability of hydro turbines, reporting that the power can increase by a factor of 3.1.
47 Regarding diffuser design, Mehmood[8] explored a number of shapes based on NACA airfoils
48 and observed a relevant improvement of velocity on the diffuser throat. However, the effect
49 of the diffuser efficiency on DAHTs is not reported.
50 The study of hydrodynamic design of DAHTs plays a very important role for improving
51 turbine technologies, since the diffuser area ratio and efficiency are important to increase the
52 power coefficient. DAHTs take advantage of the Venturi effect, reducing the fluid pressure
53 downstream and increasing the axial velocity through a contraction. The contraction is
54 located at the diffuser throat where the rotor is placed. Shrouded turbines also have the
55 advantage of allowing the turbine to start rotating at lower free stream velocities when
56 compared to the same turbine without a diffuser. These features implies a higher energy
57 production of DAHT systems [9] with a significant potential of power cost reduction [10].
58 Even though the present work deals with DAHTs, the mathematical approach comes
59 from diffuser-augmented wind turbine (DAWT) theory, and thus further works on DAWTs
60 need to be considered. For example, Shahsavarifard et al. [11] developed an experimental
61 analysis of two diffusers with different geometries, achieving maximum power coefficient of
62 0.84, far beyond the Betz-Joukowsky limit, demonstrating the potential of shrouded turbines.
63 Using both Computational Fluid Dynamics (CFD) and 1D axial momentum theory, Hansen
64 et al. [12] demonstrated that the power rise of shrouded turbines is proportional to the

2
65 increase of mass flow through the turbine blades. Rio Vaz et al. [13] developed an innovative
66 approach for the performance analysis of DAWTs based on BET, in which a more general
67 semi-empirical one-dimensional analysis is performed extending the Glauert’s correction to
68 avoid the high values of axial induction factor. This formulation is validated using CFD and
69 experimental data from Hoopen’s work [14]. Moreover, Vaz & Wood [15] have implemented
70 an algorithm to optimize the blade chord length and twist angle distributions of the turbine
71 shrouded by a diffuser, improving the rotor geometry aerodynamically. The research group
72 of Ohya performed an impressive experimental [16] and computational work [17] on both the
73 flanged [18] and lens diffuser [19]. The flange produces a higher pressure drop behind the
74 rotor, which improves the turbine performance as the recirculation increases.
75 This work proposes a new approach for the design of DAHTs using an improved BET
76 model by considering the diffuser efficiency. New expressions for the axial induction factor
77 and thrust are obtained, in order to assess DAHTs performance. A high axial induction
78 factor correction is also proposed. Comparisons are made using experimental data from the
79 literature and a CFD study for two different diffusers: Flanged Conical Diffuser (FCD) and
80 Flanged Lens Diffuser (FLD). In order to ensure accurate results, a validation using only the
81 bare turbine and an empty diffuser are carried out against benchmark experiments, NREL
82 PHASE VI [20] and Abe and Ohya flanged diffuser [17, 20], respectively. Then, the complete
83 system (turbine and diffuser coupled) is simulated. The proposed methodology exhibit good
84 agreement between experimental data and theory, presenting an error of 0.38% by comparing
85 the power output using a diffuser efficiency of 98%. Furthermore, using the CFD simulation,
86 it is observed that the wake characteristics for DAHTs is rather different when compared
87 with turbines without a diffuser, heavily influenced by the interactions between recirculation
88 zones and vortical structures behind the rotor and the diffuser. The results are compared
89 under the same conditions and a reasonable agreement for power coefficient is obtained
90 between CFD and BET. Thus, DAHT’s power coefficient can be computed accurately using
91 a lower computational resource through the proposed BET.
92 The remainder of this paper is organized as follows. The section 2 shows the geometrical
93 and operational details of the DAHT system, and presents the numerical modeling of both
94 blade element theory with a diffuser and CFD. The former subsection shows the blade
95 element theory with diffuser, in which a simple one-dimensional axial momentum theory
96 under diffuser effect is performed. This section also shows the expressions for the thrust
97 coefficient and the axial induction factor, which lead to new formulations for the performance

3
98 analysis of DAHTs. The later subsection presents the numerical modeling using CFD to
99 evaluate the proposed approach. In section 3 the numerical validation is stated, in order
100 to ensure coherent results. In section 4 the results and discussion are presented, where
101 the performance and comparisons of the proposed model is shown. Section 5 shows the
102 conclusions of this study.

103 2. Numerical Modeling

104 2.1. Geometrical and operational characteristics


105 In the simulations a 10 m diameter, 3-blade hydrokinetic turbine designed in the Energy
106 and Environment Laboratory (EEL) of the University of Brasilia is considered. The turbine
107 has a hub diameter of 1.2 m and blade shape built using NACA 653 − 618 airfoil. Main
108 turbine design specifications are summarized in Tab. 1. The airfoil’s chord length and twist
109 angle distribution along the blade are given in Figs. 1 and 2, respectively. In all DAHT
110 cases the turbine is kept with angular velocity of 20 RPM for a range of current speed (V0 )
111 from 1.5 to 3.5 m/s.

Table 1: Design parameters used in the simulation of the DAHT.


Parameters Values
Turbine Diameter (D) 10.0 m
Hub Diameter 1.2 m
Number of blades 3
Current Velocity (V0 ) 1.5 − 3.5 m/s
Water density (ρ) at 25◦ C 997 kg/m3
Angular velocity 20 RPM
Airfoil Naca 653 − 618

1.5

1
Chord [m]

0.5

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 5.5
Radial position [m]
−0.6

−0.4

−0.2

0
x [m]

0.2

0.4

0.6

0.8
0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
y [m]

Figure 1: Chord distribution.

4
0.7

0.6

Twist Angle [rad]


0.5

0.4

0.3

0.2

0.1

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 5.5
Radial position [m]

Figure 2: Twist angle distribution.

112 The diffuser assembly design consists of three parts: nozzle, diffuser and flange. The
113 nozzle is designed to smooth the flow through the diffuser. It acts as an entrance device,
114 which intends to decrease head loss due to the large scale turbulence at the inlet. Both
115 diffuser and flange sections are designed to produce pressure drop due to the strong vortex
116 formation behind it, accelerating the flow through the shrouded turbine [17].
117 Figure 3 shows the dimensional details of the two flanged diffusers: FCD and FLD. The
118 FCD has a nozzle with diameter of 1.01D, which remains cylindrical with 0.14D length,
119 presenting an opening angle of 4°. The FLD is constructed based on information extracted
120 from Hu & Wang [21]. Its main feature is the short length (0.13D) and its curved shape
121 generated from a spline revolution. Comparing both diffusers, FCD is much larger than
122 FLD. However, due to the curved shape, FLD displays a bigger area ratio.

Flanged Conical Diffuser Flanged Lens Diffuser


ø=1. 01D
0.14D

ø=1.02D
0.13D
0.5D

ø=1.09D 0.05D

area 4° area
ratio = 1. 09 ratio = 1 . 14
0.13D

Figure 3: Geometrical illustration of FCD and FLD.

5
123 2.2. Blade Element Theory with Diffuser

124 2.2.1. Axial Momentum Theory under Diffuser Effect


125 The axial momentum theory under diffuser effect considers a turbine as an ideal actuator
126 disk with an uniform velocity profile. The flow is taken as frictionless and there is no
127 tangential velocity component. To regard a diffuser with losses, an equivalent methodology
128 used in ducts [22, 23] is taken into account. As it is shown in Fig.4, the domain of interest
129 is divided into sections identified by subscripts, where 0 refers to the undisturbed region, 1
130 is just before the turbine, 2 is just after it, 3 is the diffuser outlet and 4 is far downstream.

0 1 2 3 4
V0 V0 V0 V0

V0 V1 V2 = V V3
A

A3
1

Figure 4: Control volume of an actuator disk shrouded by a FCD.

131 The actuator disk is placed between sections 1 and 2. According to Vaz & Wood [15],
132 the power coefficient in this case is given by
"  2 #
V4 2
ε2 ,

CP = ε 1 − − (1 − ηd ) 1 − β (1)
V0

133 where V0 is the freestream velocity, ε = V1 /V0 is the velocity speed-up ratio from section 0
134 to 1, β = A/A3 , where A is the rotor area (assumed equal to the diffuser area at the rotor)
135 and A3 is the diffuser exit area, and ηd is the diffuser efficiency, defined as

p 3 − p2
ηd = 1 . (2)
2
ρ (V22 − V32 )
136

6
137 As shown by Sørensen [24], the corresponding thrust may be obtained dividing the pro-
138 duced power by the local velocity at the rotor plane. Thus, the losses through the diffuser
139 are

∆Hd = p2 − p3 + 1/2ρ V22 − V32 .



(3)

140 So, applying the energy balance on the control volume depicted in Fig. 4, yields

1
V2 T = Q ρ V02 − V42 − Q∆Hd

(4)
2
141 where T is the thrust on the rotor, and Q = V2 A is the flow rate. Hence, CT becomes

 2
T V4
− (1 − ηd ) 1 − β 2 ε2 .

CT = 1 =1− (5)
2
ρAV02 V0
142 According to Rio Vaz et al. [13], CP depends on the far wake velocity, V4 , which is assumed
143 to be (1 − 2a) V0 , where a is the axial induction factor at the rotor plane, given by the
144 classical actuator disk theory. In other words, the same condition for the axial velocity in
145 the wake of an ideal bare turbine is considered in the flow after the diffuser outlet. The
146 diffuser speed-up ratio (γ = V1∗ /V0 ) is defined as the ratio between the maximum axial flow
147 velocity in the diffuser and the freestream velocity (V0 ). Thus, the velocity approaching the
148 disc becomes V1 = V2 = γ(1 − a)V0 and ε = γ(1 − a). Under these definitions, Eqs. (1) and
149 (5) result in

CP = γ (1 − a) 4a (1 − a) − γ 2 (1 − a)2 (1 − ηd ) 1 − β 2
 
(6)

150 and
CT = 4a (1 − a) − γ 2 (1 − a)2 (1 − ηd ) 1 − β 2 .

(7)

151 Clearly, if ηd = 1 in Eqs. (6) and (7), CP and CT reduce to the expressions of Rio Vaz et
152 al. [13]. Applying the energy balance as is done in the conventional actuator disk theory,
153 the thrust coefficient is obtained from CP = εCT , agreeing with the formulation found by
154 Hansen et al. [12].

155 2.2.2. A New Expression for the Axial Induction Factor


156 To include the diffuser effect into BET, the influence of the area ratio, diffuser efficiency,
157 and the speed-up ratio need to be considered by the model. Therefore, the flow angle (φ) is

7
158 defined as in Eq. (18) of Vaz & Wood [15]. The thrust coefficient is expressed as

σCn
CT = γ 2 (1 − a)2 , (8)
sin2 φ
159 where σ = Bc/2πr is the solidity. Combining Eqs. (7) and (8), the new formulation for the
160 axial induction factor yields

γ 2 σCn
 
a 2

= + (1 − ηd ) 1 − β . (9)
1−a 4 sin2 φ
161 The tangential induction factor is given by Eq. (30) of Rio Vaz et al. [13]. Note that the
162 axial induction factor, a, is dependent on ηd , γ, and β. Further, Eq. (9) reduces to the
163 one developed by Rio Vaz et al. [13] for ηd = 1. Incorporating the axial momentum into
164 BET ensures an extensible model to DAHT, allowing a practicable radial variation of the
165 diffuser speed-up ratio γ, making the approach less complex. However, it is noteworthy that
166 such variation needs to be further investigated, since it seems to modify the classical BET
167 hypothesis upon the interaction between two consecutive blade sections.

168 2.2.3. Tip Loss and High Induction Factor Corrections


169 Prandtl’s tip loss factor, F , is used in the present approach. As demonstrated by Vaz
170 & Wood [15], F is defined as the ratio of the total bound circulation of the blades and
171 the circulation of a rotor with an infinite number of blades. It is worth noting that BET
172 calculations made with the Prandtl’s factor have good agreement with free wake vortex
173 theory and experimental data [15].
174 In the literature, the axial momentum theory does not match experimental data for high
175 axial induction factors (a > 1/3), being necessary to apply corrections on the theory. Here,
176 an empirical correction for the thrust coefficient is proposed for a > 1/3, including the effect
177 of the diffuser efficiency. The first part (a < 1/3) of Eq. (10) was first proposed by Glauert
178 [25] for bare turbines. Some details on derivations and further analysis may be found in
179 Spera [26]. For a turbine with diffuser, Eq. (10) is suggested here, in which ηd and β are
180 incorporated.

1
 [4a (1 − a) − ε2 (1 − ηd ) (1 − β 2 )] F ; a≤ 3
CT = (10)
 4a 1 − a (5 − 3a) − ε2 (1 − η ) (1 − β 2 ) F ; a >
   1
4 d 3
181

182 The new formulations for a > 1/3 might be computed by combining Eqs. (8) and (10),

8
183 resulting in

K 1

1+K
; a≤ 3
a= (11)
 3a3 − (5 + 4K) a2 + (4 + 8K) a − 4K = 0; a > 1
3

184 where K accounts for the information on the diffuser efficiency and the turbine geometry,
185 which is given by

γ2
 
σCn 2

K= + (1 − ηd ) 1 − β . (12)
4 F sin2 φ
186

187 2.3. Computational Fluid Dynamics

188 The Navier-Stokes equations govern velocity and pressure fields, through which incom-
189 pressible and fully turbulent flow are considered. To take into account the turbulent phe-
190 nomena without numerically solving all eddy scales, the governing equations are solved using
191 the Reynolds Averaged Navier-Stokes (RANS) approach. Thereby, the governing equations
192 of the flow are the continuity and momentum conservation equations, as further described
193 in [27].
194 The computational domain dimensions are based on validations accurately carried out
195 by other authors [28, 29] and the guidelines developed by Rezaeiha et al. [30]. Taking those
196 studies into account, the entire domain of the present work consists of a box of 50 m x
197 50 m x 250 m as illustrated in Fig. 5. To avoid any spurious influence from the boundary
198 surfaces, the rotor is positioned at 5D and 20D from the inlet and outlet surfaces, respectively.
199 This numerical domain is also divided into two subdomains: one inner cylindrical rotational
200 volume containing the turbine and another stationary part including the surrounding. The
201 rotatory one assumes different dimensions for each case (bare and shrouded turbine). For
202 the bare turbine, the cylindrical volume is built with 10.4 m diameter and 1.2 m length, in
203 order to encapsulate the volume into the diffuser. Despite the rotating domain presenting
204 different sizes on bare and shrouded turbines, it does not cause a noticeable effect on its
205 performance [30]. The stationary subdomain is built large enough to involve the entire wake
206 region, generally divided into near and far wakes. The near wake is highly influenced by
207 the turbine hydrodynamics, having complex phenomena such as vortex shedding near the
208 blade tip [31]. It commonly extends up to 3D behind the rotor and is delimited by the point
209 that relative pressure turn equals to zero. Just after it, the far wake region, defined as the

9
210 zone of recovery of the stream velocity, may extends until 12D downstream [32]. However,
211 according to Mo et al. [33], the boundary length of near and far wake is directly influenced
212 by the Tip-Speed Ratio (T SR = ΩR/V0 ).
Free slip

pressure relative =0
velocity

5D

outlet
Inlet

Moving
Reference
Frame region

Free slip
5D 20D

Figure 5: Setup of the computational domain.

213 Since there is no consistent DAHT experimental work in the literature, the mesh config-
214 uration of all three geometries used here is based upon validation of decoupled brenchmark
215 cases of rotor [20] and diffuser [17] as shown in section 3. It is assumed that, once the mesh
216 refinement configuration has reached agreement with those experiments, it is reasonable that
217 the shrouded turbine simulations would be accurate as well. Due to the large computational
218 domain required in this work, the mesh refinement takes place in zones that the flow features
219 most affect the turbine performance: near wake and near the walls. Regarding the former,
220 the refinement covered a region from -1D up to 6D from the rotor and is set at an average
221 size of 0.15m in order to capture the pressure drop caused by the diffuser and its effect on
222 the rotor performance as is shown in Fig. 6a. The latter plays a crucial role in computing
223 the forces acting on the blade, therefore having a near wall treatment compatible with the
224 κ − ω SST turbulence model is a matter of importance. In order to achieve a high resolu-
225 tion of the boundary layer and an accurate pressure distribution on the blade, the present
226 study acknowledge the impact factors of near wall grid spacing on the NREL PHASE VI
227 highlighted by Moshfeghi et al.[34] using that turbulence model. Based upon this work, the
228 blade topology grid is built with 100 nodes per metre in the chordwise direction and 600
229 nodes along its span (Fig. 6b). In order to capture accurately the separation point, inside
230 the boundary layer, 25 prismatic nodes which grows at 1.2 spacing ratio are placed (Fig. 6c).
231 The κ − ω SST turbulence model requires lower values for y + , defined by y + = (∆yu+ /ν),
232 where ∆y is the distance of the first node from the wall, u+ is the wall shear velocity and ν
233 is the kinematic viscosity. To achieve y + < 1 the first node is positioned at ∆y = 10−6 m.

10
234 Concerning the gap between the blade and the diffuser, even tough the present study has not
235 performed a sensitivity analysis in this region, a further refinement in it may not significantly
236 affect the order of results, since the quantity of interest is the integral of forces acting on the
237 entire blade and this region just represents a small part of whole surface. Furthermore, as it
238 can be seen in Fig. 6d, this region is naturally too fine due to the proximity refinement and
239 the near wall treatment on both blade and diffuser.
240 The computational grid is made up of hybrid elements, being prisms on the near wall
241 and tetrahedrals on the rest of the domain. As shown in Tab 2, the total number of nodes
242 is about 8 × 106 and both diffuser and blade reaches y + < 1 in its maximum and average,
+ +
243 ymax and yavg respectively.

Table 2: Mesh Information.


Total Wake element Blade Diffuser
+ + + +
nodes size[m] ymax y ymax yavg
Open Turbine 8.0 ×106 0.18 0.3 0.15 - -
FCD + Turbine 7.8 ×106 0.15 0.6 0.25 0.65 0.27
FLD+ Turbine 8.0 ×106 0.15 0.25 0.1 0.54 0.21

244 A number of boundary conditions are applied in the computational domain as shown in
245 Fig. 5 and Tab. 3. First, a range of velocities is applied as a Dirichlet boundary condition
246 at the inlet. In this case, the velocity is set normal to the inlet face with 5% of turbulence
247 intensity. Second, the outlet region is also set as a Dirichlet boundary condition, but with
248 zero pressure gradient, satisfying the momentum equation. Third, the non-slip condition
249 is applied in all solid surfaces (rotor and diffuser faces). For the top, bottom and lateral
250 faces of the stationary domain, the free-slip condition is applied. Finally, the rotational
251 subdomain has a fixed angular velocity of 20 RPM. Since the RANS approach is used, the
252 interface between the rotational and stationary subdomains is made by imposing a rotating
253 Moving Reference Frame (MRF) in the flow, only inside the cylindrical subdomain. In this
254 manner, the rotor is kept in a fixed position and the governing equations are solved consid-
255 ering a modified gravitational acceleration, taking into account the Coriolis and centrifugal
256 components.

11
Table 3: Boundary Conditions.
Region Condition
Inlet 1.5 to 3.5 m/s
Outlet p=0
Rotor surface non-slip
Blade surface non-slip
Top, bottom and lateral surfaces free slip
Rotatory domain MRF
Angular velocity 20 RPM
Turbulence intensity 5%

(A)

(B)

(C) (D)

Figure 6: Mesh details: (a) wake refinement (b) blade and diffuser topology (c) near wall treatment on the
blade (d) tip gap between the blade and diffuser.

257 To perform all simulations, the commercial solver CFX is used. It is a finite volume

12
258 pressure based solver which stores the variables at the nodes. The pressure and velocity
259 decoupling is handled by modified Rie and Chow [35] discretization. In order to improve
260 its convergence this code also uses a parallelized coupled algebraic multigrid approach and
261 blending scheme which tries its utmost to use second order approximations [36]. For all
262 cases, at least 1000 iterations are performed or until the computed power has converged.
263 Here, the κ − ω SST model described by Menter et. al. [37] is employed. Due to its
264 performance in dealing with strong adverse pressure gradients and industrial flows, this model
265 has been widely used in aero and hydrodynamic studies [38]. This occurs because κ − ω SST
266 combines two well-established turbulence models, the κ − ω and κ − , in regions that they
267 fit best, near to the solid walls and free-stream flow, respectively. SST model uses auxiliary
268 switch functions, which depend on the turbulent quantities and the dimensionless normal
269 distance from the wall, y + . In this approach, κ − ω computes the turbulent contribution
270 from the inner boundary layers and κ −  calculates it out of the boundary layers, preventing
271 the excessive sensitiveness of the κ − ω model at the free-stream conditions. Consequently,
272 SST model gives great results for wall-bound flows, even in separated regions. Because of
273 these features, this model is a natural choice in flows through DAHT cases. As usual in
274 RANS calculations, since the time-averaged Navier-Stokes equations are solved, the solution
275 is computed considering steady-state regime.

276 3. Numerical Validation

277 To assure reliable and accurate numerical results, validations are performed. Due to the
278 lack of DAHT data, the methodology here is based on benchmark cases of each device solely,
279 such as NREL Phase VI turbine from National Renewable Energy Laboratory [20] (bare
280 turbine case) and the diffuser from Abe & Ohya [17] (diffuser case). Once accurate results
281 for each one of these cases is achieved, the simulation for the coupled system is considered
282 accurate as well. The computational domain dimensions and the boundary conditions are
283 set accordingly to each experimental setup and geometry.

284 3.1. NREL Phase VI turbine


285 The mesh refinement is essential to avoid incoherent converged solutions. Since it also
286 requires a large computational domain, a great refinement strategy must be done in order to
287 capture accurately all quantities. In CFD steady-state simulations, the grid size is the main
288 parameter that determines the time consuming and the resolution of the data. Then, the
289 main purpose of this mesh strategy is to refine the element small enough as its size would

13
290 not interfere in the results. In the present study, the refinement strategy focused mainly
291 in two regions: near wake and boundary layer. The near wall zone plays a key role on the
292 turbine hydrodynamic, as it presents the highest pressure and velocity gradients, as well as
293 complex phenomena, such as vortex shedding, collapse of the boundary layer and production
294 of turbulence. The near wake region, which may extend up to 3D downstream the rotor,
295 also has a great effect on the torque [39]. Thus, refining the mesh in both of these regions
296 improves the results as the torque highly depends on the boundary layer structure and the
297 evolution of turbulent scales in the wake [40].
298 To overcome this mesh arrangement issue, a grid sensitivity study is made by refining the
299 mesh in both wake and boundary layer, comparing the power with experimental data [20] for
300 u∞ = 7 m/s, as shown in Tab. 4. To reach accurate results for the computed pressure acting
301 on the blade, this work takes into account the near wall grid study as done by Moshfegui
302 et al [34]. The average element size on the blade surface is set in a 0.01 m, and at the near
303 wall 25 prismatic layers with a 1.2 growth rate are placed, where the first node is located at
304 10−6 m from the wall surface. In the near wake the average element size is kept fixed at 0.2
305 m and the refinement region increases from one mesh to the subsequent to reach acceptable
306 values of power in comparison with experimental data. On the last 3 meshes the element
307 size is decreased in the wake, in order to guarantee that any further refinement would show
308 noticeable changing regarding the computed power. As a result, Mesh 4 achieves the best
309 trade-off in terms of refinement, accuracy and computational resources. As shown in Fig. 7,
310 the pressure coefficient computed in this mesh follows the same trend as the experimental
311 data at u∞ = 7 m/s for many radial positions. Furthermore, the calculated power also
312 presents good agreement with several results in the literature for other boundary conditions
313 as depicted in Fig. 8

Table 4: Mesh sensitivity study.


Mesh 1 Mesh 2 Mesh 3 Mesh 4 Mesh 5
Total nodes 3.36 5.78 6.43 7.61 11.08
+
ymax 7.431 2.082 0.482 0.347 0.59
+
yavg 1.932 0.522 0.113 0.063 0.059
wake refinement [m] 7 14 21 21 21
element size on wake [m] 0.2 0.2 0.2 0.15 0.08
Power [kW] 3.825 4.225 5.667 6.073 6.05
relative error 36 29 5.6 1 0.8

14
(c) (a) r/R=0.3
-4
Experiment
Present work
-3

Pressure coefficient
(b) -2

-1

(a) 1

2
0 0.2 0.4 0.6 0.8 1
x/chord

(b) r/R=0.63 (c) r/R=0.95


-3 -3
Experiment
Experiment
Present work
Present work
-2 -2
Pressure coefficient

Pressure coefficient

-1 -1

0 0

1 1

2
0 0.2 0.4 0.6 0.8 1 2
0 0.2 0.4 0.6 0.8 1
x/chord
x/chord

Figure 7: Numerical and experimental data[20] of pressure coefficient at u∞ = 7m/s for three radial positions:
(a) r/R=0.3, (b) r/R=0.6, (c) r/R=0.95.

20

29
28
41
42
43

Figure 8: Power as a function of the Inlet Velocity: Comparison with results available in the literature.

314 To appraise the CFD accuracy on the wake, the numerical results are also compared with

15
315 experimental data of two sonic anemometers located at 5.84 m downstream the rotor [44],
316 as shown in Fig. 9. These probes are set 2 m one from other, being responsible to measure
317 the velocity components for a wide range of operational conditions (u∞ = 5 to 20 m/s). The
318 comparison between the computed axial velocity and the experimental data of probe #1 and
319 #2 are respectively shown in Figs. 9a and 9b, in which the abscissa indicates the velocity,
320 u∞ , set at the inlet region, and the ordinate is the axial velocity at the probe’s position. For
321 most cases, the present work (red line) achieves good agreement with the experiments (blue
322 line), as the former is predominantly on the uncertainty of the instrument measurements
323 indicated by the error bar.

Anemometer
probes
#2 #1

2m

2,45 m 5,84m

(a) (b)
25 25

20
Anemometer #1 [m/s]

20
Anemometer #2 [m/s]

Present work
15 44
Experimental [22] 15

Present work
Experimental [22]
44
10 10

5 5

0 0
0 5 10 15 20 25 30 0 5 10 15 20 25 30
u [m/s]

u [m/s]

Figure 9: Comparison of axial velocity between experimental [44] (blue line) and numerical (red line) on:
(a) Anemometer #1 (b) Anemometer #2.
.

324 3.2. Diffuser Validation


325 For the diffuser cases, the convergence of results are assessed using three mesh levels, and
326 then the selected grid is compared with the literature. For the FCD, the convergence for the
327 axial normalized velocity along the center line is clearly seen (Fig. 10a). The main discrep-
328 ancy occurs at the peak in which the coarsest and finest mesh differ about 3%. Comparing
329 with experimental data [17] (Fig. 10b), despite differences of 5% found at the peak, overall
330 the selected grid also is able to capture the trend of the velocity curve. The Venturi effect

16
331 is clearly noted at X = 0.4D, as the local velocity increases 1.6 times the current velocity
332 (u0 ). Even though the experimental data along the axial coordinate direction is not very
333 large, it can be expected that the velocity would be fully recovered at X = 5D downstream
334 diffuser. For the FLD case, multiple mesh refinements are used and the convergence is also
335 achieved as shown in Fig. 11a. However, in this case, the results are compared against
336 another numerical work [21] which shows the same curve trend of axial normalized velocity
337 (Fig. 11b). In the FLD case, the peak of the speed-up ratio is about 1.2 at X = 0.1D and
338 it still affects the flow up to 1D downstream.
1.8 1.8
Mesh 0.97x10 6 nodes Abe and Ohya [18](Experimental)
1.7 1.7
Mesh 4.3x10 6 nodes Mesh 6x10 6 nodes
1.6 Mesh 6x106 nodes 1.6
1.5 1.5
1.4 1.4

U/Uo
U/Uo

1.3 1.3

1.2 1.2

1.1 1.1

1 1

0.9 0.9

0.8 0.8
-2 -1 0 1 2 3 4 -2 -1 0 1 2 3 4
(a) X/D (b) X/D

Figure 10: Normalized Velocity on the axial direction on FCD case: (a) convergence (b) validation.

1.3 1.3
Mesh 0.7M nodes
Mesh 8.7M nodes Present work
1.25 1.25 CFD data [21]
Mesh 11.7M nodes

1.2 1.2

1.15 1.15
U/U0
U/U0

1.1 1.1

1.05 1.05

1 1

0.95 0.95

0.9 0.9
-2 -1.5 -1 -0.5 0 0.5 1 1.5 2 2.5 3 -2 -1 0 1 2 3
(a) X/D (b) X/D

Figure 11: Normalized Velocity on the axial direction on FLD case: (a) Convergence (b) comparison with
other CFD simulation [21].

17
339 4. Results and Discussion

340 4.1. Performance of the Improved BET model

341 According to Laws & Epps [45] lift-based, axial-flow turbines use the same principles
342 as air-craft wings, propellers, and wind turbines. In this case, the blades of a lift-based
343 turbine are composed of two-dimensional hydrofoil cross-sections, as previously described
344 along the section 2.2. Therefore, due to the difficulties in finding works with diffuser detailed
345 geometry, thrust and power coefficient data for DAHTs in the literature, the use of data
346 available for wind turbine theory becomes acceptable, even because all design parameters
347 are dimensionless. Therefore, in this section, the improved BET model is compared using
348 four major references. The first one is the classical experimental data obtained by Moriarty
349 & Hansen [46] for a bare turbine, whose data for thrust coefficient are depicted in Fig. 12a.
350 The second one is the CFD modeling developed by Hansen et al. [12], in which an actuator
351 disc CFD model of the flow through a wind turbine in a diffuser is developed. The results are
352 shown in Fig. 12b and Tab. 5. The third reference is the result obtained by Rio Vaz et al.
353 [13], whose data are also shown in Fig. 12b and Tab. 5. The last one, is the measurements
354 developed by Hoopen [14], which are presented in Tab. 7.
355 Hence, to analyze the influence of the diffuser efficiency, ηd , on the power and thrust
356 coefficients, CP and CT , respectively, Eqs. (6) and (7) are used. For CT , the results are
357 generated by varying ηd and considering an area ratio of β = 0.278. This value was chosen
358 only for the purpose of evaluating the model behavior. As raised in Barbosa et. al. [47], a
359 large area ratio may cause flow separation within the diffuser, consequently changing the flow
360 through the rotor. Similar assumptions are also done in [18]. Figure 12a shows the results
361 obtained using the proposed high induction factor correction. For a diffuser of maximum
362 efficiency (ηd = 1), the turbine behaves as a bare turbine, since CT → 4a(1 − a) in Eq.
363 (7). Note that as much as CT increases the thrust becomes gradually more like at of a
364 bare turbine. This result is important because it demonstrates the heavy influence of ηd on
365 CT , agreeing with those reported by Hansen et al. [12] and Phillips [23], which suggest that
366 increasing CT must decrease the circulation created by the diffuser, leading to a flow behavior
367 through the rotor similar to a bare turbine. Additionally, increasing energy losses along the
368 diffuser by friction action, the axial induction factor tends to increase as well, leading to a
369 decreased diffuser speed-up ratio, ε = γ(1 − a). Such features are also observed by Phillips
370 [23], through which the present work has showed a good physical behavior, converging to
371 the experimental data when ηd = 1.

18
372 Figure 12b shows the CP in relation to CT . To evaluate the impact of ηd on CP , the
373 numerical results developed by Hansen et al. [12] is used. In their work, ηd = 0.83, γ = 1.83,
374 and β = 0.54 are taken into account. As shown in Tab. 5, the proposed analysis has good
375 agreement with the CFD [12] modeling. The present work reaches about 5.3% error for
376 CP opt , while Rio Vaz et al. [13] achieves 14.8%. This result shows that the proposed model
377 is more accurate because the model account the diffuser efficiency. Such optimum power
378 coefficient is reached for CT opt = 0.79 using the present model which is indeed closer to the
379 one calculated by Hansen et al. [12], in which the relative error is only of 1.25%.

2 1.5
Experimental Without difuser (Theoretical)
ηd = 0.3 Rio Vaz et al. [13]
= 0.6 Present work (Theoretical)
1.5 CFD (Hansen et al. [12 ])
= 1.0
1

P
CT

1
C
Betz-Joukowsky limit
0.5
0.5

0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
a C
T
(a) (b)
Figure 12: (a) Correction for the thrust coefficient (experiment obtained from [46]). (b) Effect of ηd on CP .

Table 5: Comparison between the present work and CFD [12].


Hansen et al. [12] Rio Vaz et al. [13] Present work
CT opt 0.80 0.89 0.79
CP opt 0.94 1.08 0.89

380 All the results shown so far demonstrated only the theoretical behavior of the improved
381 BET model regarding the impact of the diffuser efficiency. Now, to analyze the accuracy
382 of the proposed model and to produce evidence of simulation reliability, the real turbine
383 experienced by Hoopen [14] is employed. Thus, a three-bladed wind turbine with 1.5 m
384 diameter rotor, for an undisturbed stream velocity of 10 m/s is simulated. The diffuser is a
385 circular airfoil with β equal to 0.5785. The blades and diffuser are designed and optimized
386 using the results computed by CFD [14]. The output plane of the diffuser has a diameter
387 of 2 m and is equipped with a Gurney flap of 0.04 m. Hoopen [14], reported a torque equal
388 to 7.10 Nm for a wind speed of 10 m/s and Ω = 75 rad/s (≈ 716.2 rpm). The Reynolds

19
389 number based on the diffuser mean diameter is 6 × 105 . Delta-shaped vortex generators are
390 installed on the diffuser trailing edge in order to promote mixing. The diffuser speed-up
391 ratio is shown in Fig. 13. The DAWT geometry and the experimental parameters of the
392 diffuser speed-up ratio are summarized in Tab. 6 and Fig. 13, respectively.

Table 6: Design parameters used in this work.


Parameters Values
Turbine Diameter 1.5 m
Hub Diameter 0.3 m
Number of blades 3
Diffuser airfoil NLR [14]
Blade airfoil NACA 2207
Power output 531 W
Freestream Velocity (V0 ) 10 m/s
Air density (ρ) at 20◦ C 1.2 kg/m3
Rotational speed 716.2 rpm

0.8
Radial position, r/R

0.6

0.4

0.2

0
0.8 0.9 1 1.1 1.2 1.3 1.4
Diffuser speed-up ratio, γ(r/R)

Figure 13: Diffuser speed-up ratio as a function of the radial position [14].

393 Table 7 presents the power output and torque for V0 = 10 m/s, where the present work
394 is compared with experimental data. Using the improved BET, the result is more accurate
395 than that made by Rio Vaz et al. [13]. This occurs because into BET formulation the
396 present model uses the contribution of the area ratio, β, and the pressure loss due to friction
397 through the parameter ηd . The use of those two parameters is extremely relevant in the
398 design of turbines with a diffuser. As followed by Philips [23] through experimental results
399 the performance of a turbine surrounded by diffuser strongly depends on ηd and β. This
400 observation is also described by Sørensen [24], which also shows how the area ratio may
401 impact the maximum power coefficient of a turbine with a diffuser.

20
402 Therefore, the relative error obtained using the modified BET is of 0.38% for the power
403 output at Ω = 75 rad/s, for ηd = 98%. The CFD results of Hoopen [14] used Ω = 137
404 rad/s, which is much larger than the experimental one. The diffuser airfoil and details of
405 the modeling can be found in [14]. To show the main contributions of this paper, Tab. 8
406 depicts the differences between the proposed formulations and that described by Rio Vaz et
407 al. [13]. The main modifications occurred in the axial induction factor and in the thrust
408 coefficient for introducing the effect of ηd and β, whose modifications are responsible for the
409 good behavior of the model, as well as bringing an important tool for the current state of
410 the art.

Table 7: Comparison between the present work and experimental data (V0 = 10m/s).
Angular speed Power output Torque
(rad/s) (W) (Nm)
Experimental [14] 75 531 7.10
Rio Vaz et al. [13] 75 526 6.10
Present work 75 533 7.12
CFD [14] 137 545 4.00
CFD [14] 155 246 1.60

Table 8: Main difference between the present work and Rio Vaz et al. [13].
Axial induction Thrust coefficient
factor (a) (CT )
Rio Vaz et al. [13] a
h 1−a
= γ 2 BcC n
2
8πr sin φ i
C T = 4a(1 − a)
2
Present work a
1−a
γ BcC n
2
2
= 4 2πr sin φ + (1 − ηd ) (1 − β ) CT = 4a (1 − a) − γ (1 − a)2 (1 − ηd ) (1 − β 2 )
2

411 4.2. Comparisons between CFD Simulation and the Improved BET model

412 The mass flow increment caused by the diffuser is shown in Tab. 9. The mass flow
413 rate for the turbine with FCD is about 20% higher than the bare one, while the turbine
414 using FLD is increased only 2.4%. This result suggests that FCD has better performance
415 compared with FLD.

Table 9: Mass flow rate at the rotor plane.

Device Mass flow rate ( u0ṁρA )


FCD 1.29
FLD 1.23
Turbine 0.82
FCD + Turbine 0.99
FLD + Turbine 0.84

416 In order to assess the increment of the flow velocity on the rotor, a projection of its area
417 located at 0.1D before the rotor plane is used as a reference (Fig. 14) . Higher values of the

21
418 axial normalized velocity (uy /U∞ ) are achieved for FCD. On both types, maximum mass flow
419 rates are located at dimensionless radial positions, r/R, between 0.8 and 1. Furthermore,
420 FCD simulations also experience instabilities in the boundary layer on the outer section.
421 Once the turbine is placed inside the diffusers, the rotational direction of the forthcoming
422 blade affects uy /U∞ radially covering a region beyond the blade projection. Overall, as shown
423 in Tab. 10, FCD presents both the highest value (1.26) and larger average (0.83) of axial
424 normalized velocity. The turbine shrouded by the lens diffuser presents a slight increment
425 on these quantities when compared to the open system.

u /U
y 0 Flanged Lens Diffuser (FLD) Flanged Conical Diffuser (FCD)

Turbine + FLD Turbine + FCD


Open turbine

Figure 14: Axial normalized velocity at 0.1D from the rotor plane on three configurations: diffuser augmented
turbines, open turbine and only the diffusers.

Table 10: Axial normalized velocity at different configurations.


γ
highest average
Bare 1.09 0.7
FCD 1.26 0.83
FLD 1.10 0.75

426 Figure 15 shows the behavior of the axial normalized velocity (uy /U∞ ), pressure and
427 Turbulence Kinetic Energy (TKE) at TSR=5.6, in order to assess how the wake is influenced
428 by the diffuser augmented turbine. Concerning the pressure curve shown in Fig 15a, all cases
429 presented a similar pattern. Pressure increases strongly just before the rotor and as it passes

22
430 through it, an abrupt drop is experienced by the flow. The pressure is only recovered about
431 5D downward. Overall, the diffuser does not affect much the pressure recovery along the
432 axial direction. In Fig 15b it can be seen that the velocity recovery presents a significant
433 disparities in each system. In general, the velocity falls dramatically due to the adverse
434 pressure gradient at the rotor position and then experiences a slow steady rise up to regain
435 its initial value. Once this data is collected at the centerline, the negatives values and
436 fluctuations up to 2.5D are mainly caused by the nacelle and root vortices shed. While on
437 the open turbine it takes up to 15D to be fully recovered, on shrouded ones may extends
438 until 20D downstream the rotor. Figure 15c depicts that the turbulence intensity (TI) rise
439 heavily experiencing its peak just after the rotor due to stall on the root blade. As it moves
440 downward, the TI plunges and it goes up again. This second increasing is mainly due to the
441 mixture of tip and root vortices at the centerline. This subsequent decay indicates that the
442 spiralling vortices from the tip and root has collapsed and broken down into smaller scales
443 eddies [33]. It can be noted that the diffuser promotes a forehand TI rise and smooth its fall
444 over the domain. As a result on both DAHT systems the TI is still significant up to 20D
445 downward the turbine.

23
4000
Open Turbine
3000 FLD+Turbine
FCD+Turbine
2000

1000
Pressure [Pa]
0

−1000

−2000

−3000

−4000
−4 −2 0 2 4 6 8 10 12 14 16 18 20
2 Y/D
Open Turbine
FLD+Turbine
FCD+Turbine
Axial Normalized Velocity [u /U ]
0

1.5
y

0.5

−0.5
−4 −2 0 2 4 6 8 10 12 14 16 18 20
Y/D
0.8
Open Turbine
0.7 FLD+Turbine
FCD+Turbine
0.6
Turbulence Intensity [%]

0.5

0.4

0.3

0.2

0.1

0
−4 −2 0 2 4 6 8 10 12 14 16 18 20
Y/D

Figure 15: Axial normalized velocity and static pressure along the wake at T SR = 5.6.

446 To assess the impact of the diffusers on the power output, the extended BET model
447 is compared with the CFD modeling. The diffuser speed-up ratio, γ, as a function of the
448 rotor radial position is shown in Fig. 16, which is necessary to BET in order to calculate
449 the turbine performance. Despite both diffusers FCD and FLD achieve similar increment
450 of velocity at the blade tip, the FCD presents higher values of γ on a larger radial section.

24
451 In this case, the diffuser efficiencies are 83% for the FCD, and 16% for the FLD, calculated
452 using the averages of pressures and velocities through Eq. 2.
453 Figure 17a shows that the power coefficient for the DAHT with FCD increases 69% when
454 compared with the bare turbine at TSR = 5.17. The standard deviation of the differences
455 between the CFD simulation and the present work are of 0.0257 for the turbine without
456 diffuser, 0.0242 for the turbine with FCD, and 0.0429 for the turbine with FLD. Despite
457 the disparities between CFD and BET, the calculations of power coefficient present good
458 agreement for the turbine with FCD, and reasonable for the other ones. It is believed that the
459 main discrepancies found between CFD and BET on DAHT with FLD are strongly related
460 to the flow complexities such as hydrodynamics interactions between tip vortex and other
461 turbulent structures and the boundary layer flow at the diffuser walls, high pressure drop
462 and unsteadiness, which are not considered on a simplistic analysis as BET. In addition,
463 the maximum value of Cp occurs for TSR = 5.17, when the turbine is surrounded by FCD.
464 It may happen due to the increasing velocity near the rotor plane as previously shown in
465 Tab. 10. Figure 17b shows an increased power output of 53% when compared to the bare
466 turbine, suggesting FCD as a good diffuser in terms of energy generation. Even tough, the
467 power coefficient using FCD shown in Fig. 17a does not exceed the Betz-Joukowsky limit,
468 the diffuser effect increases the turbine efficiency 1.7 times when compared with the bare
469 turbine. This result shows that a diffuser needs to be carefully dimensioned to recover as
470 much kinetic energy as possible from the flow, so that depending of the diffuser technology,
471 a DAHT may not exceed the Betz-Joukowsky limit.

1.2

0.8
γ(r/R)

0.6

0.4

0.2 FLD
FCD
0
0 0.5 1 1.5
Radial position, r/R

Figure 16: Diffuser speed-up ratio, γ, as a function of the radial position, r/R.

25
0.7 600
Betz-Joukowsky limit BET - no diffuser
0.6 CFD - no diffuser
500 BET - FCD
0.5 CFD - FCD
400 BET - FLD

P - (kW)
0.4 CFD - FLD
Cp

300
0.3 BET - no diffuser
CFD - no diffuser 200
0.2 BET - FCD
CFD - FCD
100
0.1 BET - FLD
CFD - FLD
0 0
3 4 5 6 7 1.5 2 2.5 3 3.5
TSR V0 - (m/s)

(a) (b)
Figure 17: (a) Power coefficient vs. TSR for open and shrouded turbines. (b) Power output as functions of
the freestream velocity.

472 5. Conclusions

473 The following major conclusions are derived from the present study:

474 ˆ New formulations to the axial induction factor and thrust coefficient applied to Diffuser
475 Augmented Hydrokinetic Turbines are presented.

476 ˆ The model has demonstrated to be more accurate than that developed by Rio Vaz et
477 al. [13], yielding good agreement with CFD and experimental data.

478 ˆ The approach is an important tool for computing thrust and power coefficients as
479 functions of the angular velocities for DAHT.

480 ˆ This new approach is based on BET, allowing computational cost several times lower
481 than complete CFD simulations.

482 ˆ The main model’s limitation is the necessity of a prediction of the diffuser speed-up
483 ratio, γ, preferably as an explicit function of the radius, which can be done by a
484 CFD simulation of the diffuser solely, since it is much cheaper computationally than a
485 complete DAHT system.

486 ˆ The new model assumes the hypothesis proposed by Glauert where the far wake stream-
487 velocity V4 is equal to (1 − 2a)V0 , which still need to be verified by detailed wake
488 numerical simulations or even experimental investigations.

26
489 ˆ The cavitation also need to be considered in the design of DAHTs, since the diffuser
490 effect must increase such a phenomenon.

491 ˆ Despite such limitations, the results obtained in this work present a consistent be-
492 havior. The results show that diffuser technologies lead to a relevant improvement of
493 the DAHT’s power output, demonstrating that such a technology is very important to
494 hydrokinetic turbine design.

27
Nomenclature
Abbreviations
BET Blade Element Theory
CF D Computational Fluid Dynamics
EEL Energy Environment Laboratory
F CD Flanged Conical Diffuser
F LD Flanged Lens Diffuser
N REL National Renewable Energy Laboratory
RAN S Reynolds Averaged Navier Stokes
SST Shear Stress Transport
Arabic Symbols
a, a0 Axial and tangential induction factors
A, A3 Swept area of turbine blades, and exit area of the diffuser (m2 )
B Number of blades
c Chord (m)
cp3 Diffuser exit pressure
CP Power coefficient
Cn , Ct Normal and tangential force coefficients
CT Thrust coefficient
D Diffuser diameter at the inlet (m)
Di Drag force (N/m)
fj Force per unit of volume(N/m3 )
Fn , Ft Normal and tangential forces (N/m)
h Height of flange (m)
K Parameter defined by Eq. (12), which carries information on the effect
of the diffuser and the rotor geometry
L Diffuser length (m)
Li Lift force (N/m)
mb mass flow for a bare wind turbine (kg/s)
md mass flow for a wind turbine with diffuser (kg/s)
p0 , p2 , p3 Pressure in ambient free-stream and pressure at rotor diffuser
and exit-plane, respectively (Pa)
P Power output (W)
p Averaged pressure (Pa)
r, R Radial position and turbine radius (m)
Re Reynolds number
Rec Local Reynolds number based on the airfoil chord
T Thrust force (N)
ui Averaged velocity components (m/s)
0 0
ui uj Reynolds Stress tensor
Vi Flow velocity (subscript 0 stands for free-stream, 1 and 2 for the rotor plane,
3 for the diffuser exit-plane, and 4 for downstream of the turbine)(m/s)
V1∗ Maximum axial velocity of flow in the diffuser (m/s)
Xturbine Position where the turbine should be installed (m)
w, W Angular wind velocity and relative velocity (m/s)

28
Greek Symbols
α Angle of attack
γ Diffuser speed-up ratio
ηd Diffuser efficiency
β Exit-area ratio
 Velocity speed-up ratio
θ Twist angle
ρ Air density (kg/m3 )
σ Solidity
Ω Angular velocity of the rotor (rad/s)
µ Viscosity (Pa.s)
∆Hd Pressure loss through the diffuser [Pa]

495 6. Acknowledgments

496 The authors would like to thank the CNPq, PROPESP/UFPA, LEA-UnB, and ELETRONORTE
497 for financial support.

498 [1] N. Kolekar, A. Banerjee, Performance characterization and placement of a marine hy-
499 drokinetic turbine in a tidal channel under boundary proximity and blockage effects,
500 Applied Energy 148 (2015) 121–133.

501 [2] W. Schleicher, J. Riglin, A. Oztekin, Numerical characterization of a preliminary


502 portable micro-hydrokinetic turbine rotor design, Renewable Energy 76 (2015) 234–241.

503 [3] C. Belloni, R. Willden, G. Houlsby, An investigation of ducted and open-centre tidal
504 turbines employing cfd-embedded bem, Renewable Energy 108 (2017) 622–634.

505 [4] A. Bahaj, A. Molland, J. Chaplin, W. Batten, Power and thrust measurements of marine
506 current turbines under various hydrodynamic flow conditions in a cavitation tunnel and
507 a towing tank, Renewable energy 32 (3) (2007) 407–426.

508 [5] P. Mycek, B. Gaurier, G. Germain, G. Pinon, E. Rivoalen, Experimental study of the
509 turbulence intensity effects on marine current turbines behaviour. part i: One single
510 turbine, Renewable Energy 66 (2014) 729–746.

511 [6] P. A. S. F. Silva, L. D. Shinomiya, T. F. de Oliveira, J. R. P. Vaz, A. L. A.


512 Mesquita, A. C. P. B. Junior, Analysis of cavitation for the optimized de-
513 sign of hydrokinetic turbines using bem, Applied Energy 185 (2016) 1281–1291.
514 doi:http://dx.doi.org/10.1016/j.apenergy.2016.02.098.

29
515 [7] D. Gaden, E. Bibeau, A numerical investigation into the effect of diffusers on the per-
516 formance of hydro kinetic turbines using a validated momentum source turbine model,
517 Renewable Energy 35 (2010) 1152–1158.

518 [8] N. Mehmood, L. Zhang, J. Khan, Exploring the effect of length and angle on naca 0010
519 for diffuser design in tidal current turbines, in: Applied Mechanics and Materials, Vol.
520 201, Trans Tech Publ, 2012, pp. 438–441.

521 [9] V. Ramos, G. Iglesias, Performance assessment of tidal stream turbines: a parametric
522 approach, Energy conversion and management 69 (2013) 49–57.

523 [10] A. Vazquez, G. Iglesias, Device interactions in reducing the cost of tidal stream energy,
524 Energy Conversion and Management 97 (2015) 428–438.

525 [11] M. Shahsavarifard, E. Bibeau, A. Birjandi, Performance analysis of open and ducted
526 wind turbines, IEEE 978-0-933957-40-4 2013 MTS.

527 [12] M. O. L. Hansen, N. N. Sørensen, R. Flay, Effect of placing a diffuser around a wind
528 turbine, Wind Energy 3 (4) (2000) 207–213.

529 [13] D. A. T. D. Rio Vaz, A. L. A. Mesquita, J. R. P. Vaz, C. J. C. Blanco, J. T. Pinho, An


530 extension of the blade element momentum method applied to diffuser augmented wind
531 turbines, Energy Conversion and Management 87 (2014) 1116–1123.

532 [14] P. Hoopen, An experimental and computational investigation of a diffuser augmented


533 wind turbine with an application of vortex generators on the diffuser trailing edge, M.Sc.
534 Thesis. Faculty of Aerospace Engineering, Delft University of Technology.

535 [15] J. R. Vaz, D. H. Wood, Aerodynamic optimization of the blades of diffuser-augmented


536 wind turbines, Energy Conversion and Management 123 (2016) 35–45.

537 [16] Y. Ohya, T. Karasudani, A. Sakurai, K. Abe, M. Inoue, Development of a shrouded wind
538 turbine with a flanged diffuser, Journal of wind engineering and industrial aerodynamics
539 96 (5) (2008) 524–539.

540 [17] K. Abe, M. Nishida, A. Sakurai, Y. Ohya, H. Kihara, E. Wada, K. Sato, Experimental
541 and numerical investigations of flow fields behind a small wind turbine with a flanged
542 diffuser, Journal of Wind Engineering and Industrial Aerodynamics 93 (12) (2005) 951–
543 970.

30
544 [18] K. Abe, Y. Ohya, An investigation of flow fields around flanged diffusers using cfd,
545 Journal of wind engineering and industrial aerodynamics 92 (3) (2004) 315–330.

546 [19] Y. Ohya, T. Karasudani, A shrouded wind turbine generating high output power with
547 wind-lens technology, Energies 3 (4) (2010) 634–649.

548 [20] M. M. Hand, D. Simms, L. Fingersh, D. Jager, J. Cotrell, S. Schreck, S. Larwood, Un-
549 steady aerodynamics experiment phase VI: wind tunnel test configurations and available
550 data campaigns, National Renewable Energy Laboratory Golden, Colorado, USA, 2001.

551 [21] J.-F. Hu, W.-X. Wang, Upgrading a shrouded wind turbine with a self-adaptive flanged
552 diffuser, Energies 8 (6) (2015) 5319–5337.

553 [22] K. Foreman, B. Gilbert, R. Oman, Diffuser augmentation of wind turbines, Solar Energy
554 20 (4) (1978) 305–311.

555 [23] D. G. Phillips, et al., An investigation on diffuser augmented wind turbine design, Ph.D.
556 thesis, ResearchSpace@ Auckland (2003).

557 [24] J. Sørensen, Wind Turbine Technology: fundamental concepts of wind turbine engineer-
558 ing, Springer International Publishing, 2016. doi:10.1007/978-3-319-22114-4.

559 [25] H. Glauert, Aerodynamic theory. in: Durand wf, editor. chapter xi. division l. airplanes
560 propellers, reprinted, Dover, NewYork, 1963 4 (1935) 191–195.

561 [26] D. Spera, Wind Turbine Technology: fundamental concepts of wind turbine engineering,
562 ASME Press, 2nd Ed., New York, 2009.

563 [27] C. Argyropoulos, N. Markatos, Recent advances on the numerical modelling of turbulent
564 flows, Applied Mathematical Modelling 39 (2) (2015) 693–732.

565 [28] J.-O. Mo, Y.-H. Lee, Cfd investigation on the aerodynamic characteristics of a small-
566 sized wind turbine of nrel phase vi operating with a stall-regulated method, Journal of
567 mechanical science and technology 26 (1) (2012) 81–92.

568 [29] N. N. Sørensen, J. Michelsen, S. Schreck, Navier–stokes predictions of the nrel phase vi
569 rotor in the nasa ames 80 ft× 120 ft wind tunnel, Wind Energy 5 (2-3) (2002) 151–169.

570 [30] A. Rezaeiha, I. Kalkman, B. Blocken, Cfd simulation of a vertical axis wind turbine
571 operating at a moderate tip speed ratio: guidelines for minimum domain size and az-
572 imuthal increment, Renewable Energy 107 (2017) 373–385.

31
573 [31] P. A. Silva, T. F. OLIVEIRA, A. C. Brasil Junior, J. R. Vaz, Numerical study of wake
574 characteristics in a horizontal-axis hydrokinetic turbine, Anais da Academia Brasileira
575 de Ciências (AHEAD) (2016) 0–0.

576 [32] B. Sanderse, S. Pijl, B. Koren, Review of computational fluid dynamics for wind turbine
577 wake aerodynamics, Wind energy 14 (7) (2011) 799–819.

578 [33] J.-O. Mo, A. Choudhry, M. Arjomandi, R. Kelso, Y.-H. Lee, Effects of wind speed
579 changes on wake instability of a wind turbine in a virtual wind tunnel using large eddy
580 simulation, Journal of Wind Engineering and Industrial Aerodynamics 117 (2013) 38–
581 56.

582 [34] M. Moshfeghi, Y. J. Song, Y. H. Xie, Effects of near-wall grid spacing on sst-k-ω model
583 using nrel phase vi horizontal axis wind turbine, Journal of Wind Engineering and
584 Industrial Aerodynamics 107 (2012) 94–105.

585 [35] C. Rhie, W. L. Chow, Numerical study of the turbulent flow past an airfoil with trailing
586 edge separation, AIAA journal 21 (11) (1983) 1525–1532.

587 [36] C. Ansys, Solver theory guide, Ansys CFX Release 11 (2006) 1996–2006.

588 [37] F. R. Menter, Zonal two equation k-turbulence models for aerodynamic flows, AIAA
589 paper 2906 (1993) 1993.

590 [38] F. Menter, M. Kuntz, R. Langtry, Ten years of industrial experience with the sst tur-
591 bulence model, Turbulence, heat and mass transfer 4 (1) (2003) 625–632.

592 [39] M. Moshfegh, Y. H. Xie, Effects of near-wall grid spacing on sst-k-omega model using
593 nrel phase vi horizontal axis wind turbine, Journal of wind Engineering and Industrial
594 Aerodynamics.

595 [40] L. Chamorro, M. Guala, R. Arndt, F. Sotiropoulos, On the evolution of turbulent scales
596 in the wake of a wind turbine model, Journal of Turbulence (13) (2012) N27.

597 [41] M. A. Potsdam, D. J. Mavriplis, Unstructured mesh cfd aerodynamic analysis of the
598 nrel phase vi rotor, AIAA paper 1221 (2009) 2009.

599 [42] J.-C. Huang, H. Lin, T.-J. Hsieh, T.-Y. Hsieh, Parallel preconditioned weno scheme for
600 three-dimensional flow simulation of nrel phase vi rotor, Computers & Fluids 45 (1)
601 (2011) 276–282.

32
602 [43] A. L. Pape, J. Lecanu, 3d navier–stokes computations of a stall-regulated wind turbine,
603 Wind Energy 7 (4) (2004) 309–324.

604 [44] S. Larwood, Wind turbine wake measurements in the operating region of a tail vane,
605 National Renewable Energy Laboratory.

606 [45] N. Laws, B. Epps, Hydrokinetic energy conversion: Technology, research, and outlook,
607 Renewable and Sustainable Energy Reviews 57 (2016) 1245–1259.

608 [46] P. Moriarty, A. Hansen, Aerodyn theory manual. tech. rep. nrel/tp 500-36881, Tech.
609 rep., Golden, CO: National Renewable Energy Laboratory (2005).

610 [47] D. Barbosa, J. Vaz, S. Figueiredo, M. Silva, E. Lins, A. Mesquita, An investigation of a


611 mathematical model for the internal velocity profile of conical diffusers applied to dawts,
612 Annals of the Brazilian Academy of Sciences 87 (2). doi:http://dx.doi.org/10.1590/0001-
613 3765201520140114.

33

You might also like