Sie sind auf Seite 1von 13

Etienne Gauthier

Laboratoire de Mecanique des


Fluides Numerique,
Department of Mechanical Engineering,
Laval University,
Quebec City, QC G1V 0A6, Canada
e-mail: etienne.gauthier.4@ulaval.ca

Thomas Kinsey
Laboratoire de Mecanique des
Fluides Numerique,
Department of Mechanical Engineering,
Laval University,
Quebec City, QC G1V 0A6, Canada
e-mail: thomas.kinsey.1@ulaval.ca

Guy Dumas1
Laboratoire de Mecanique des
Fluides Numerique,
Department of Mechanical Engineering,
Laval University,
Quebec City, QC G1V 0A6, Canada
e-mail: gdumas@gmc.ulaval.ca

Impact of Blockage on the


Hydrodynamic Performance
of Oscillating-Foils
Hydrokinetic Turbines
This paper describes a study of the impact of confinement on the hydrodynamic performance of oscillating-foils hydrokinetic turbines (OFHT). This work aims to contribute to
the development of standards applying to marine energy converters. These blockage
effects have indeed to be taken into account when comparing measurements obtained in
flumes, towing tanks, and natural sites. This paper provides appropriate correction formula to do so for OFHT based on computational fluid dynamics (CFD) simulations performed at a Reynolds Number Re 3  106 for reduced frequencies between f * 0.08
and f * 0.22 considering area-based blockage ratios ranging from e 0.2% to 60%.
The need to discriminate between the vertical and horizontal confinement and the impact
of the foil position in the channel are also investigated and are shown to be of secondorder as compared to the overall blockage level. As expected, it is confirmed that the
power extracted by the OFHT increases with the blockage level. It is further observed
that for blockage ratio of less than e 40%, the power extracted scales linearly with e.
The approach proposed to correlate the performance of the OFHT in different blockage
conditions uses the correction proposed by Barnsley and Wellicome and assumes a linear
relation between the power extracted and the blockage. This technique is shown to be
accurate for most of the practical operating conditions for blockage ratios up to 50%.
[DOI: 10.1115/1.4033298]

Introduction

With highly energetic and predictable resources, tidal and river


flows offer an energy potential that could greatly help the transition toward a cleaner global production of electricity. However,
this resource is not yet exploited and there is still an important
need for research to accelerate its development and to support the
establishment of standards for the industry. In the development of
hydrokinetic turbines, it becomes essential to be aware of the differences between the design and operating conditions of these systems. As full-scale testing is not financially feasible at preliminary
steps, model testing allows to validate the performance of a device
prior to deployment. These models tested in water canals are
affected by the proximity of walls constraining the flow. Natural
sites may also be characterized by various levels of blockage,
from local bathymetry, shallow waters, relative proximity with
seabed and free surface. Proper correction factors for blockage
effects will allow more accurate comparisons of performance
measurements between various laboratory facilities and natural
sites.
An important knowledge base on confinement effects has been
derived from wind tunnel testing. Research on wall blockage
effects has been conducted since the 1930s, starting with Glauert
[1] for propeller testing. The Glauert correction, based on an onedimensional momentum approach, is still widely used in testing to
correct for wind tunnel interference.
Maskell [2] developed a theory for blockage correction of bluff
bodies in closed-section wind tunnel based on the momentum balance in the bypass flow region. This correction was the first correction which applies to the models with separated flows such as
stalled wings. Similarly, Pope and Harper [3] have provided rules
for blockage correction of different bluff bodies for wind tunnel
testing. However, it should be noted that for bluff bodies, all the
1
Corresponding author.
Contributed by the Fluids Engineering Division of ASME for publication in the
JOURNAL OF FLUIDS ENGINEERING. Manuscript received June 22, 2015; final manuscript
received March 10, 2016; published online May 26, 2016. Assoc. Editor: Elias Balaras.

Journal of Fluids Engineering

flows bypass the effective aerodynamic body which is not the case
when considering the flow through a turbine. These corrections
for bluff bodies are thus not expected to accurately account for
blockage effect in the specific case of hydrokinetic turbines.
More recently, Bahaj et al. [4] performed experimental measurements on a three-bladed horizontal-axis hydrokinetic turbine
(HAHT) in a cavitation tunnel and a towing tank. They corrected
their results to account for blockage using a method based on an
actuator disk model of a hypothetical turbine which was first proposed by Barnsley and Wellicome [5] (hereafter referred to as the
BW correction). In the latter, the Glauerts equations have been
modified to account for the wake expansion of the turbine. The
BW correction depends on the blockage ratio e, the velocity in the
testing facility U, the drag coefficient of the turbine CD, and the
estimated velocity through the turbine a2U which is calculated
iteratively. This technique is used in the present study to correlate
the results of the OFHT in different confined environments and
details are presented in Appendix A.
Sorensen et al. [6] proposed a blockage correction based on the
same analysis but which necessitates measurement of the wake
width instead of the drag coefficient, doing so the correction is
obtained explicitly.
Blockage effects on the performance of tidal turbines have also
been investigated recently in order to propose modifications to the
classical LanchesterBetz limit [7,8] which formally applies for
turbines considered unconfined. Such an adaptation to this theory
has been proposed by Garret and Cummins [9] for cases presenting a certain level of blockage in a closed-section channel (no
free-surface).
Houlsby et al. [10] followed the same procedure as Garret and
Cummins by studying the streamtube associated with the flow
through a hypothetical turbine but with an extension to open channel flow. They presented a step by step procedure to compute the
theoretical power available from an unconfined case (equivalent
to the LanchesterBetz limit), a confined flow by horizontal
planes and an open-channel flow with confinement through freesurface proximity. The free-surface case has also been studied by

C 2016 by ASME
Copyright V

SEPTEMBER 2016, Vol. 138 / 091103-1

Downloaded From: http://asmedigitalcollection.asme.org/ on 08/07/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Whelan et al. [11]. In the present study, we focus on the effect of


confinement from horizontal planes.
In the case of cross-flow turbines such as OFHT and verticalaxis hydrokinetic turbines (VAHT), the unsteady nature of the
dynamics of these systems presents an additional challenge to
evaluate blockage effects on performance. For OFHT, the impact
of the channel blockage has never been explored until now. In the
development of this novel hydrokinetic turbine technology, most
simulations have been performed considering an unconfined environment (e.g., see Refs. [1220]). To effectively do so in the cited
works, the authors increased the distance between the domain
boundaries and the turbine until no variation on the hydrodynamics was observed.
This paper thus presents the first ever results of a study on the
impact of confined flow conditions on the hydrodynamic performance of OFHT. The main objective here is to propose a proper
blockage correction approach to extrapolate the performance of
OFHT in different flow confinements. For that purpose, we have
selected the simplest possible configuration in order to investigate
the essence of confinement effects on the active part of the turbine, the oscillating foil in the present case. It is left to future
investigations to incorporate the effects of the presence of the supporting structure and linkage components whose design is likely
to differ among turbine models.

heaving amplitude is H0/c 1, and / is kept constant at 90 deg.


The reduced frequency is defined as f * fc/U which is varied
from f * 0.08 to f * 0.22 in this study. It should be noted that
the selection of these parameters is the result of previous investigations carried out by the authors group at Laval University
[1217]. The optimal performance of the turbine is expected to be
achieved within this range of parameters for a plausible largescale deployment.
Based on a recent study [17] performed at Reynolds number of
500,000 with a NACA0015, it was found that dynamic stall was
not necessary to obtain optimal performance in power extraction
(CP 43%). Since a thicker hydrofoil is characterized by a larger
leading-edge radius which helps to delay dynamic stall while providing an improved mechanical strength, a NACA0025 hydrofoil
is considered in the present study. As will be presented later in
Fig. 7, it allows to reach performance up to CP 47% in the
unconfined scenarios.
As indicated earlier, the simulations in this paper are performed
at a chord Reynolds number Re 3  106 which is quite representative of actual river and tidal applications as can be inferred from
Table 1. This Reynolds number being well above the typical transitional regime, we assume here fully turbulent boundary layers
and thus avoid the uncertainties associated with the
laminarturbulent transition modeling.
Forces and moment coefficients are used to analyze the performance and are defined as

Problem Definition
CX t

2.1 Oscillating Hydrofoils. The present work considers an


NACA0025 hydrofoil of aspect ratio b/c 10 at a Reynolds number (Re Uc/) of Re 3  106 undergoing a combined heaving
h(t) and pitching h(t) motion in a uniform flow as sketched in
Fig. 1. The position of the foil is defined as
ht h0 sinc t

CY t

(1)
CM t

ht H0 sinc t /

(2)

and the corresponding velocities


Xt h0 c cosc t

(3)

Vy t H0 c cosc t /

(4)

where h0 and H0 are, respectively, the pitching and heaving amplitude, X is the pitching velocity, Vy is the heaving velocity, c is the
angular frequency (2pf), and / is the phase difference between
the two motions. In this study, the pitching axis is located on the
chord line at a distance xp/c 0.4 from the leading edge, c being
the chord length. The pitching amplitude is h0 80 deg, the

Xt
1
q U2 b c
2
Y t
1
q U2 b c
2
Mt
1
q U 2 b c2
2

(5)

(6)

(7)

where X is the instantaneous horizontal hydrodynamic force, Y is


the instantaneous vertical hydrodynamic force, M is the instantaneous torque at the pitching axis, and b is the foil span. The total
power extracted by the hydrofoil (P) is the sum of the power contributions associated to the heaving (Py) and to the pitching (Ph)
motions which are defined as
Py t Yt Vy t

(8)

Ph t Mt Xt

(9)

Pt Py t Ph t

(10)

To evaluate the performance of the system in a similar way to


what is made for VAHT and HAHT, the mean power extracted is
compared to the upstream power available Pa through the swept
area of the turbine (AT HTWT 2 H0b)
CP 

Py Ph
P

Pa 1
q U 3 AT
2

(11)

Table 1 Typical range of parameters for OFHT targeted


applications

Fig. 1 Imposed heaving h(t) and pitching h(t) motions.


Adapted from Ref. [12].

091103-2 / Vol. 138, SEPTEMBER 2016

Chord length (m)


Free stream velocity (m/s)
Power (kW)
Reynolds number ( 106)

Tidal

River

13
23.5
804000
1.58.8

0.52
1.53
81000
0.65

Transactions of the ASME

Downloaded From: http://asmedigitalcollection.asme.org/ on 08/07/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

This normalized power is limited to a maximum theoretical


value of 16/27 or 59% by the LanchesterBetz limit, which
applies to a single idealized turbine in an unconfined domain. Let
us recall that this limit is derived from the analysis of an inviscid
stationary stream tube [7,8]. The drag coefficient related to the
mean CX is defined for convenience as
CD

X
c
CX
1
2 H0
2
q U AT
2

(12)

Blockage Parameters

Figure 2 presents the principal parameters to define the blockage ratio of a turbine in a channel. Following the Nomenclature
usually adopted in aero and hydrodynamics [1,3,4,9], we define
the blockage ratio e as the ratio of the turbine swept area AT to the
cross-sectional area of the channel A

AT WT HT

A
W H

(13)

where WT, HT, W, and H are the frontal dimensions of the turbine
and the domain as presented in Fig. 2. The confinement can be analyzed in terms of vertical confinement HT/H and horizontal confinement WT/W. A turbine is considered centered in the channel
when HT/H  WT/W. The impact of vertical and horizontal confinements is addressed in Sec. 6.1. Moreover, the values of blockage ratio are limited by e 100% for a fully confined turbine and
e 0% for an unconfined one.

Numerical Methodology

The finite volume code CD-AdapcoTM STAR-CCMV V9 is


used in this work with a segregated flow solver [21]. The segregated flow model can be described as using a collocated variables
arrangement and a Rhie-and-Chow-type pressurevelocity coupling combined with a semi-implicit method for pressure-linked
equations algorithm (SIMPLE). Second-order schemes are used
for pressure, momentum, and turbulent quantities discretizations.
The unsteady formulation is based on a second-order implicit
scheme.
The turbulence modeling uses an unsteady Reynolds-Averaged
NavierStokes (RANS) approach with the kx shear-stress transport (SST) turbulence model [22] with curvature correction [21].
This model has been chosen for its versatility and its recognized
reliability for applications with and without important flow separations. Furthermore, several comparisons with previous studies
on oscillating hydrofoil turbines by the authors group [12,1416]
have shown that the predictions of the kx SST model agree well
with those using the SpalartAllmaras model or the scale-adaptive
simulation model especially as far as mean power and instantaneous force coefficients are concerned.
A no-slip wall boundary condition is imposed on the hydrofoil
where the boundary layers are meshed in order to ensure a
y  O1. All simulations were run in time until statistical convergence was reached, i.e., a variation of the mean power coefficient CP from one cycle to another of less than 0.2%. Further,
within each time step, iterative convergence was monitored
through instantaneous forces. Typically, monitored values stabilized in about five iterations.
The boundary conditions and the main dimensions of the computational domain are presented in Fig. 2. The channel length is
fixed at L 80 c and the channel width W and height H are modified to obtain different blockage ratios. In this study, the turbine
height and width are, respectively, HT 2 c and WT 10 c. The
symmetry condition allows the simulation of only half the geometry and therefore decreases the computational cost. A symmetry

Fig. 2 Computational
conditions

domain

and

exterior

boundary

condition which is equivalent to a slip-wall boundary condition is


enforced on both lateral boundaries as well as on the top and bottom boundaries. For all the simulations, the turbulence settings at
the inlet are a turbulence intensity of 0.1% and a turbulence viscosity ratio of 0.01, i.e., essentially an unperturbed quiet upstream
flow. Higher levels of inlet turbulent intensity, more representative of natural flow conditions, are not expected to affect significantly the present fully turbulent simulations nor the relative
impact of blockage.

Journal of Fluids Engineering

4.1 Meshing Approach. The overset mesh approach, also


known as chimera grid, is used to take into account the motion of
the foil while solving in a stationary frame of reference. This technique allows to simulate bodies in relative motions without having
to use other techniques such as deforming mesh or remeshing. It
also tolerates very close proximity motions which may be the case
for high blockage ratio where the hydrofoil moves near the channel boundaries. As presented in Fig. 3, the overset mesh and background mesh are superposed. When the simulation is initialized,
the cells of the background mesh in the superposed zone are deactivated. The regular discretized equations are solved in the active
cells and interpolation is made within the layer of acceptor cells.
For the background grid, a structured mesh type is used. The
zone in which the foil oscillates and its near wake is discretized
with isotropic cells with a bunching toward the foil tip. Figure 4
presents a vertical slice at the midspan of the foil showing the
mesh characteristics for the e 50% case.
4.2 Time and Space Refinements. The quality of our time
and space discretizations was ascertained in order to ensure an
independence of the results regarding the grid refinement and the
time step size. The 50% blockage ratio case has been used to
study that effect. Table 2 presents the mesh characteristics for
three levels of discretization. Mesh refinement is made here by
varying the number of nodes on the foil section as well as in the
spanwise direction (Dz).
For the time discretization, 500, 1000, and 2000 time steps per
cycle were tested with the three-dimensional (3D)-3 mesh to isolate the impact of time resolution. As shown in Table 3, maximum
variations of 1.4% were observed on force and moment coefficients between 1000 and 2000 time steps per cycle. Therefore, a
1000 time steps per cycle discretization has been chosen for this
study. To evaluate the impact of space refinement, the results of
our three grids are compared using 1000 time steps per cycle.
SEPTEMBER 2016, Vol. 138 / 091103-3

Downloaded From: http://asmedigitalcollection.asme.org/ on 08/07/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 3 Description of the overset mesh technique and the interface characteristics

Fig. 4 Characteristics of the mesh shown on a vertical slice at a midspan position for a
blockage of e 5 50%

Differences of less than 1% on the forces and moment coefficients


are observed between meshes 3D-2 and 3D-3. The spatial discretization of the 3D-2 mesh is thus shown to be adequate and has
been used for this study.

Validation

To demonstrate the reliability of the present computational


method, two series of comparisons with experimental data are
proposed.
First, the drag polar of an NACA 0025 hydrofoil with aspect
ratio of 6 (square tips foil) has been computed using the present
numerical approach with the same level of spatial resolution as
the mesh 3D-2 (Table 3 and Fig. 4) and the same turbulence
model and boundary conditions. The results from a series of nine
stationary RANS simulations at different angles of attack
(12 deg < a < 18 deg) are shown in Fig. 5. The agreement with
the wind tunnel data of Bullivant [23] is quite satisfactory. It is
found that lift is accurately predicted within 2% (root-meansquare difference over the whole range of angles) while drag is
slightly over predicted. At this Reynolds number of 3  106, it is
expected and reasonable that the present fully turbulent RANS
simulations (without transition model) predict slightly higher friction drags than the measured ones on the free transition foil of the
experiments.
For the second series of comparisons, we reproduce an oscillating foil turbine case that is well documented experimentally and
numerically. It involves an NACA 0015 hydrofoil of aspect ratio
7 (with endplates described in Ref. [16]) with a pitching axis at
Table 2 Meshes characteristics
Mesh

Total cells ( 106)

Nodes on foil section

Dz

3D-1
3D-2
3D-3

4.65
9.44
24.71

200
292
400

0.04 c
0.03 c
0.02 c

091103-4 / Vol. 138, SEPTEMBER 2016

x c/3 and a Reynolds number of 500,000. Using the exact same


approach as described in Sec. 4 together with a similar mesh resolution as the present 3D-1 mesh and the basic time discretization
of 1000 timesteps per cycle, we obtain the results shown in Fig. 6
for the power coefficient. A total of six values of reduced frequency have been computed among which five are the same as the
ones simulated in Refs. [14,16] with an entirely different flow
solver (ANSYS-FLUENT) and methodology (general grid interface,
heaving reference frame, SpalartAllmaras turbulence model).
The lower frequency cases (f * 0.080.12) are characterized
with important dynamic stall and significant leading edge vortex
shedding (LEVS) occurring on a significant portion of the cycle.
The higher frequency cases (f * 0.140.20), on the other hand,
are free of LEVS in the midstroke where most of the power is
extracted. For those latter cases, the agreement between the present simulations and the previous independent simulations of Ref.
[16] is excellent. Both sets of numerical predictions tend, however, to overestimate slightly the experimental data which remains
quite satisfactory considering that this latter value is in fact a minimal bound estimate [24]. For the three lower frequencies, the
present predictions are within 3% of the experimental data which
is again quite satisfactory. The fact that they are slightly below
Table 3 Time and space refinements for the case e 5 50%,
cycle-averaged values of CX, and peak values of CY and CM are
provided as well as the cycle-averaged power coefficient CP
Mesh

ts/cycle

CX

C^Y

C^M

CP

Time
3D-3
3D-3
3D-3

500
1 000
2 000

2.525
2.509
2.474

3.462
3.553
3.525

0.950
0.952
0.952

0.775
0.791
0.784

Space
3D-1
3D-2
3D-3

1 000
1 000
1 000

2.467
2.490
2.509

3.494
3.522
3.553

0.947
0.954
0.952

0.784
0.789
0.791

Transactions of the ASME

Downloaded From: http://asmedigitalcollection.asme.org/ on 08/07/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 5 Hydrodynamic polar of a NACA 0025 rectangular wing


of aspect ratio 6 at a Reynolds number Re 5 3.0 3 106 (experimental data from Bullivant [23])

Fig. 7 Dependency of the cycle-averaged power coefficient CP


on blockage for different reduced frequencies f*

the previous numerical predictions of Ref. [14] seems attributable


to the differences in the turbulence modeling within the LEVS
region. Nonetheless, the general trend predicted by the present
simulations and the global agreement between them and both the
previous simulations and the experimental data is deemed very
good and adequate for the present purpose of studying the blockage effects on the turbines performance.

at arctanVy t=U  ht

Results

The dynamics of the oscillating foil depends, among other


things, on the frequency at which it oscillates. To evaluate the
impact of blockage effects in different operating conditions, simulations are performed at four different reduced frequencies
(f * 0.08, 0.12, 0.18, and 0.22). Figure 7 shows the variation of
the mean power coefficient CP for blockage ratios of e 0.2%,
12.4%, 24.8%, and 49.7%.
One can observe that the bell-shaped curves (presence of a
maximum) are present as expected. This trend can be explained
by a drop of power extracted at low frequencies because of
dynamic stall and at high frequencies by smaller effective angles
of attack leading to lower vertical forces on the foil. Two examples of flow fields are shown in Fig. 8, one presenting dynamic
stall and another associated to a small maximum effective angle
of attack. The latter is defined as

Fig. 6 Power coefficient (CP ) with respect to the reduced frequency for an unconfined NACA 0015 oscillating foil of aspect
ratio 7 at Re 5 500,000 (xp/c 5 0.33, H0/c 5 1.0, h0 5 75 deg)

Journal of Fluids Engineering

(14)

and its maximum value through the cycle can be approximated


[13] by evaluating the value at the quarter cycle t T/4
amax  ja T=4 j jarctanc H0 =U  h0 j
jarctan2 p f  H0 =c  h0 j

(15)

Figure 9 shows that even for a fixed reduced frequency,


f * 0.12 in this case, the blockage effects change the flow dynamics. One can observe at t 3 T/8 that for a blockage ratio of
e 50%, the flow separation is more important than in the unconfined case. This is attributed to a greater effective angle of attack
perceived by the foil in the confined case. Indeed, an increased
blockage leads to an increased free-stream velocity which
involves a greater effective angle-of-attack (see Eq. (14)). For
these operating conditions, at a blockage of e 50%, the foil is on
the edge of dynamic stall but still yields a very high performance
as can be seen in Fig. 7.
As shown in Fig. 7, going from f * 0.12 to f * 0.08, the
strong drop in performance is associated to the occurrence of
dynamic stall. Lowering the reduced frequency leads to a higher
maximum effective angle of attack which tends to increase the
vertical force on the foil and therefore increases the power
extracted. However, when this angle reaches a critical value leading to flow separation, an important drop of performance is
observed and the power extracted becomes essentially independent of the blockage ratio. This is the case with a reduced frequency of f * 0.08 as seen in Fig. 7.
This relative independence of power extracted with respect to
blockage ratio at f * 0.08 is due to two canceling effects. On the
one hand, the instantaneous force amplitude increases with blockage (increased dynamic pressure available), but the perceived
angle of attack also increases with blockage leading to a hastened
flow separation on the plunging foil, which leads to a worsened
synchronization between the instantaneous vertical force and velocity as shown in Fig. 10. This synchronization is important [13]
because the product of the instantaneous vertical force and the
heaving velocity gives the power extracted from the heaving
motion, Eq. (8).
For a specific reduced frequency of f * 0.18, the evolution of
power extracted as a function of the blockage ratio is presented in
Fig. 11. For blockage ratios of less than 40%, the power extracted
seems to scale linearly with the blockage ratio. However, if blockage ratios of 50% and 60% are taken into account, the general
trend seems better fitted with an exponential function.
SEPTEMBER 2016, Vol. 138 / 091103-5

Downloaded From: http://asmedigitalcollection.asme.org/ on 08/07/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 8 Evolution of vorticity fields (red counterclockwise and blue clockwise) for an unconfined hydrofoil (e 5 0.2%) at reduced frequencies of f * 5 0.08 and 0.22

Fig. 9 Evolution of vorticity fields (red counterclockwise and blue clockwise) at a reduced
frequency of f * 5 0.12 for an unconfined hydrofoil (e 5 0.2%) and for a blockage of e 5 49.7%

It should be noted that the 1.43 slope observed for f * 0.18 in


Fig. 11 is not a universal result of oscillating-foils turbine because
this slope changes depending on the operating condition (reduced
frequency) as presented in Fig. 12. It is also observed that for
lower reduced frequencies (f * 0.08), dynamic stall occurs and
the power extracted becomes independent of the blockage.
Excepting cases presenting dynamic stall, it seems reasonable,
from these results, to assume a linear relationship between the
power extracted and moderate levels of blockage ratio (e < 40%).
This assumption could potentially be used in the development of a
blockage correction formula for OFHT. It should be noted that the
drag coefficient CD also presents a linear behavior with e.
6.1 Horizontal Versus Vertical Confinements. Table 4
presents the OFHT results at f * 0.18 for different horizontal and
vertical confinements. It is observed that for both vertical and horizontal confinements, the power extracted increases with the
blockage ratio.
091103-6 / Vol. 138, SEPTEMBER 2016

The need to discriminate between horizontal and vertical confinements can be studied by comparing the cases presenting a
same blockage ratio but different confinement asymmetry (CA)
defined as
HT =H
(16)
CA
WT =W
A turbine is considered perfectly centered when CA 1. The
different channel cross sections tested are presented in Fig. 13.
Results in Table 4 show that for high blockage ratios, e.g.,
e 33%, the performance seems to be unaffected by the aspect
ratio of the channel. However, when the CA of two channels are
very different, the impact can be significant as observed for the
e 12% cases with a relative difference of about 6% on the power
extracted. Figure 14 presents the instantaneous power coefficient
signal for a different CA for a fixed blockage ratio of e 12%. It
is shown that the more the channel is narrow (WT/W ! 1) the
more it gets close to the two-dimensional (2D) result (corresponding to the case CA e).
Transactions of the ASME

Downloaded From: http://asmedigitalcollection.asme.org/ on 08/07/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 10 Evolution of instantaneous vertical force coefficient CY


at f * 5 0.08 for e 5 0.2% and e 5 49.7% along with the instantaneous heaving velocity VY /U

Fig. 12 Mean power coefficient CP normalized by the mean


power coefficient of the unconfined case (CP e 5 0.2%) with
respect to blockage ratio for different reduced frequencies. Linear regression curves are shown for 0% e 50%.
Table 4 OFHT results at a reduced frequency of f * 5 0.18 presenting different vertical and horizontal confinements
e (%)
0.2
5.0
5.0
12.4 (2D)
12.4
12.4
12.4
12.4
24.8
24.8
33.1
33.1
40.9
49.7
60.5

Fig. 11 Mean power coefficient CP normalized by the mean


power coefficient of the unconfined case (CP e 5 0.2%) with
respect to blockage ratio for a reduced frequency of f * 5 0.18. A
linear regression curve for 0% e 40% and an exponential
regression curve for 0% e 60% are also presented.

In light of these results, for CA of practical water channels or


towing tanks between 0.3 < CA < 3, there is no need to discriminate
the impact of vertical and horizontal confinements. Also, for blockage ratios of less than e 5%, the discrimination of vertical and horizontal confinements is not relevant anymore because of the great
distance between the turbine and the channel boundaries. However,
caution is advised for extreme cases in which WT/W ! 1 or HT/H
! 1, the effects of vertical and horizontal confinements might have
to be discriminated especially for moderate confinement.

6.2 Effect of Asymmetry. Depending on the deployment


site, the turbine can be fixed from the top, the bottom, or even the
sides. The turbine is not expected to be perfectly centered in the
channel, it is thus important to evaluate the performance sensitivity of an off-centered turbine. To do so, two different cases sharing the same blockage ratio e 12% and frequency f * 0.18 are
tested with two levels of vertical offset as presented in Fig. 15.
Results demonstrate that the vertical position of the turbine in
the channel has a weak impact on performance. For the case
Journal of Fluids Engineering

WT/W

HT/H

CA

CP =CP e0:2%

100
20
44.7
1
11.5
17.5
28.4
46
11.5
23
11.5
17.25
11.5
11.5
10.5

100
20
8.9
16.1
14
9.1
5.7
3.5
7
3.5
5.25
3.5
4.3
3.5
3.2

0.10
0.50
0.22
1.0
0.87
0.57
0.35
0.22
0.87
0.43
0.87
0.58
0.87
0.87
0.95

0.02
0.10
0.22
0.124
0.14
0.22
0.35
0.57
0.29
0.57
0.38
0.57
0.47
0.57
0.63

0.20
0.20
1.00
0.124
0.16
0.39
1.00
2.63
0.33
1.31
0.44
0.99
0.54
0.66
0.67

1.00
1.09
1.09
1.43
1.25
1.21
1.18
1.17
1.38
1.35
1.48
1.47
1.59
1.78
2.13

shifted from 1/4 H, a variation of 0.2% is observed on the mean


power coefficient relative to the centered case. Even for an
extreme case where the foil oscillates close to the bottom boundary (3/8 H), the impact remains as little as 1%.
Figure 16 presents a comparison of the instantaneous power
coefficients and mean streamwise velocity profiles between the
three cases. It is shown that at an upstream distance of x 2 c
(turbine being at x 0), the flow is more slowed down as the foil
comes closer to the bottom boundary. Since the foil is no more in
the middle of the channel, the velocity profile becomes asymmetric and differences are observed on the peak values of the power
coefficient. In fact, the peak values of the instantaneous power
coefficient are higher during the upstroke (t/T 0.75) than during
the downstroke (t/T 0.25) where the foil is going down toward a
zone of lower velocity. Offsetting the turbine position downward
in the channel leads to an accelerated flow in the upper part of the
channel and a decelerated flow in the lower part of the channel
when compared to the symmetric arrangement. This asymmetry
leads to small variations in the instantaneous effective angle-ofattack perceived by the foil when the latter is moving up (slightly
higher a) and down (slightly reduced a). However, these variations are small and the overall impact of this parameter can be
neglected when the interest is on cycle-averaged performance values and not the details of the instantaneous forces and power
SEPTEMBER 2016, Vol. 138 / 091103-7

Downloaded From: http://asmedigitalcollection.asme.org/ on 08/07/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 13 Dimensions of the channel cross sections tested

The BW correction (Barnsley and Wellicome [5]) used and


popularized by Bahaj et al. [4] is based on the conservation of
mass, energy, and momentum on a control volume around an actuator disk. This approach is described in more details in
Appendix A. In the latter, equations are expressed in terms of
ratios involving the wake velocity deficit a4. A hypothetical
unconfined reality yielding the same flow velocity through the turbine and the same drag than those of the confined case is considered (noted with a 0 sign). Manipulating equations and using an
iterative process, the ratio U=U 0 is obtained (see Eq. (A24)) which
represents the ratio of the upstream velocity of the confined turbine U to the equivalent upstream velocity of the hypothetical
unconfined turbine U 0 .
Knowing that the power extracted varies with the cube of
velocity and that the drag varies with velocity squared, the BW
correction ultimately implies the following expressions involving
the velocity ratio of Eq. (A24)

Fig. 14 Power coefficient evolution for different channel CA


and the limit 2D case for e 5 12.4% and f * 5 0.18

C0P CP

 3
U
U0

(17)

C0D CD

 2
U
U0

(18)

where C0P and C0D are, respectively, the equivalent power and drag
coefficients of the corresponding hypothetical unconfined turbine
while CP and CD are the power and drag coefficients of the confined turbine. Similarly, the reduced frequency can be correlated
for different blockage ratios with the following expression:
f0 f

Fig. 15 Off-centered turbines

evolutions. It should be noted that these conclusions are valid for


a closed-section channel and that free-surface effects is yet to be
evaluated.

Blockage Corrections

The blockage correction is a way to convert the results obtained


in a given channel test site in order to predict the performance of
the same turbine operating in a different flow confinement.
091103-8 / Vol. 138, SEPTEMBER 2016

 
U
U0

(19)

where f 0 and f * are, respectively, the equivalent reduced frequency of the corresponding hypothetical unconfined turbine and
the reduced frequency of the confined turbine. For a turbine with
known optimal operating conditions, this relation could allow to
estimate the reduced frequency corresponding to the peak performance if the turbine was installed in another confined environment. Equation (15) for the maximum effective angle of attack
returns the same result regardless of the blockage. However, by
using f 0 instead of f * in this expression, the maximum effective
angle of attack perceived by the confined foil (a0max ) is obtained.
The latter is more representative of the actual flow dynamics as it
takes into account the blockage.
The BW correction is applied to the present OFHT results for
the power and drag coefficients as presented in Fig. 17. Ideally,
the correction formula should bring all the points to fit on a single
curve.
Transactions of the ASME

Downloaded From: http://asmedigitalcollection.asme.org/ on 08/07/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 16 Impact of an off-centered turbine on its performance for a blockage ratio of e 5 12% and a reduced frequency of f * 5 0.18. (a) Mean streamwise velocity profiles at an upstream distance of x 5 2 c from the turbine. (b)
Evolution of the instantaneous power coefficient on the oscillating cycle showing the differences on peak values
between the downstroke and the upstroke.

Comparing the uncorrected results to the ones corrected with


the BW correction, the curves tend to overlap. Despite this good
overall trend, maximum relative difference of 15% are observed
on the C0P at low effective reduced frequency (f 0 < 0:10) where
dynamic stall occurs for our OFHT (see, e.g., Fig. 8). However,

for more common frequencies, between f 0 0:10 and f 0 0:22,


the maximum relative difference remains less than 7%. For the
drag coefficient values C0D , the maximum relative difference is
limited to 5% except again for cases presenting dynamic stall
(f 0 < 0:10).

Fig. 17 Blockage correction of Barnsley and Wellicome applied to the present OFHT results CP (f*; e) and CD (f*; e)
0
0
and the corresponding hypothetical unconfined results CP f 0 ; e and CD f 0 ; e

Journal of Fluids Engineering

SEPTEMBER 2016, Vol. 138 / 091103-9

Downloaded From: http://asmedigitalcollection.asme.org/ on 08/07/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 18 Procedure to correlate the performance of an OFHT for


different levels of blockage

The BW correction based on the performance and the drag of a


confined turbine allows to find the performance of a corresponding hypothetical unconfined turbine. Knowing the performance of
both confined and unconfined turbine and assuming the linear
relation between the power extracted and the blockage ratio, it is
possible to predict the performance of the turbine in other blockage conditions. Thus, the correction put forward by Barnsley and
Wellicome combined with the linear relation between the power
extracted and the blockage ratio represents a promising technique
to correlate the performance of OFHT in different confined flow
environments. Figure 18 illustrates this procedure further.
It should be noted that the same procedure may be used to correlate the CD for different blockage ratios. Using this technique
with the results at f * 0.18, it is shown that the relative error
between the predictions and the actual CFD results is less than 7%
on both CP and CD . This observation is limited to blockage ratios
e  50% and excludes cases presenting dynamic stall. It is important to note that the reduced frequency is also affected by this procedure. More explicitly, we have the following relations:
 

CP e1  CP0

e CP0
(20)
CP e
e1


CD e1  CD0
CD e
e CD0
(21)
e1
 

f e1  f 0
f  e
e f0
(22)
e1

Conclusion

In this study, it was shown, as expected, that the power extracted


by the OFHT increases with the blockage level. Moreover, it was
observed that for blockage ratio of less than 40%, the relation
between the power extracted and the blockage ratio scales linearly
when operating conditions do not involve massive dynamic stall.
This study also confirms that the blockage ratio simply defined
with the swept area is sufficient to characterize confinement and
the relative performance of a given turbine once the performance
at one level of confinement (CP ; CD ) is known.
The level of blockage is expected to have a significant impact
on the performance of a turbine in a confined environment but
other parameters can come into play. The need to discriminate
091103-10 / Vol. 138, SEPTEMBER 2016

between the vertical and horizontal confinement and the impact of


the position of the OFHT in the channel has been investigated for
a closed section channel. It was found that for moderate CA
(0.3 < CA < 3), there is no need to discriminate the horizontal and
vertical confinements. This parameter turns out to be of second
order as compared to the overall blockage ratio e. Also, for blockage ratios of less than e 5%, the discrimination of vertical and
horizontal confinements is unnecessary because of the large distance between the turbine and the channel boundaries. However,
it may remain necessary to discriminate vertical and horizontal
confinements for extreme cases where WT/W ! 1 or HT/H ! 1.
For the position of the turbine in the channel, it was shown that
even for an extreme case where the foil oscillates close to the
channel boundaries, variations of less than 1% on the mean power
are observed.
Furthermore, it was observed that for off-design operating conditions for which important dynamic stall occurs (f * < 0.10), performance becomes essentially independent of the blockage ratio.
The blockage correction developed by Barnsley and Wellicome
[5] was applied to the OFHT results. The desired trend and collapse of the data have been obtained and the maximum relative
spreading is limited to 7% for C0P and 5% for C0D considering frequencies between f 0 0:10 and f 0 0:22.
For high Reynolds number conditions, the performance of the
present single-foil hydrokinetic turbine in various confined flow
environments can be predicted within 7% using the BW correction
and assuming a linear evolution of the power extracted (or drag)
with the blockage ratio.
Future work should investigate the impact of confinement from
free-surface on the OFHT performance. The study presented in
this paper is part of a broader project which aims to develop general blockage corrections that could be applied to the different
types of hydrokinetic turbines (OFHT, HAHT, and VAHT).

Acknowledgment
Financial support from NSERC Canada, FRQNT Quebec, Marine Renewable Canada and Laval University are gratefully
acknowledged. A special thanks to Mike Shives from the University of Victoria as well as Robert J. Cavagnaro from the University of Washington for their useful inputs on blockage effects.
Computations were performed on the supercomputers Colosse at
Laval University and Guillimin in Montreal, both managed by
Calcul Quebec and Compute Canada. This essential resource is
here gratefully acknowledged.

Nomenclature
A
AT
b
c
CP
CP
C0P
CD
C0D
CA
f
f*
f0
g
h
H
HT
H0
L

channel cross-sectional area


turbine swept area, 2bH0
foil span, 10 c in this study
chord length
cycle-averaged power coefficient, normalized with the foil
surface, P=1=2 q U 3 b c
cycle-averaged power coefficient, normalized with the
turbine swept area, P=1=2 q U3 AT
CP relative to an equivalent unconfined environment
cycle-averaged drag coefficient, X=1=2 q U2 AT
CD relative to an equivalent unconfined environment
confinement asymmetry
frequency of oscillation of period T, 1/T
nondimensional frequency, fc/U
f * relative to an equivalent unconfined environment
gravitational acceleration
instantaneous vertical position of the hydrofoil pitching
axis
height of the channel
height of the turbine swept area
heaving amplitude
channel length
Transactions of the ASME

Downloaded From: http://asmedigitalcollection.asme.org/ on 08/07/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

M
P
Pa
Py
Ph
Re
t
T
U
U0
Vy
W
WT
X
xp
Y
a
aT=4
c
e
h
h0
q
/
X

^

moment at the pitching axis xp


instantaneous total power extracted, Py Ph
total power available, 12 q U 3 AT
heaving contribution to instantaneous total power, Y Vy
pitching contribution to instantaneous total power (MX)
Reynolds number, Uc/
time
oscillation period, 1/f
free stream velocity
U relative to an equivalent unconfined environment
instantaneous heaving velocity
width of the channel
width of the turbine swept area
horizontal component of hydrodynamic force
chordwise position of pitching axis
vertical component of hydrodynamic force
effective angle of attack, arctanVy =U  h
effective angle of attack at quarter-period ( ja T=4 j  amax )
an, bnvelocity ratios (n 2, 4 in Fig. 19)
angular frequency, 2pf
blockage ratio, WTHT/WH
instantaneous angular position of the airfoil chord w/r to
horizontal
pitching amplitude
fluid density
phase shift between pitching and heaving motions
pitching angular velocity
mean value over one motion cycle (cycle-averaged
quantity)
peak value over one motion cycle

where b4 is the ratio of the bypass velocity at station 4 to the


upstream velocity (station 1).

Bernoulli
Using Bernoulli for stations 1 to 4, 1 to 2, and 3 to 4, leads to
the following expressions:


1
p  p4 q U 2 b24  1
2


1
p2  p3 q U 2 b24  a 24
2

(A3)
(A4)

Momentum Budget
The momentum budget on a control volume including the whole
channel gives a first expression for the drag on the turbine


(A5)
D q U 2 A  A  A4 b24  A4 a 24 p  p4 A
Substituting Eq. (A3) in Eq. (A5), we get
 

 1



D q U 2 A4 b24  a 24  A b24  1 q A U2 b24  1
2
(A6)
which can be expressed in terms of A4 only using continuity
1
D qA4 U 2 b4  a4 b4 2a4  1
2

(A7)

Appendix A: Blockage Correction


Pressure Discontinuity Across the Turbine
Continuity
With respect to Fig. 19, continuity equation for the stream tube
is defined with the velocity ratios an and bn as
A1 At a 2 A4 a 4

(A1)

where a2 is the ratio of the streamtube velocity at station 2 to the


upstream velocity (station 1) and a4 is the ratio of the streamtube
velocity at station 4 to the upstream velocity (station 1). The continuity equation for the whole channel is defined as
A  A4 b 4 A4 a 4 A

(A2)

Another expression for the drag on the turbine can be derived


from the pressure difference on both side of the turbine
D At p2  p3

(A8)

and by substituting Eq. (A4) in Eq. (A8), we get




1
D q At U2 b24  a 24
2

(A9)

We have now two equations for the drag on the turbine that
must be equal. Equaling Eq. (A7) and Eq. (A9) and using continuity, we get

Fig. 19 Sketch of the flow past an idealized hydrokinetic turbine showing the expansion of
the stream tube and the different velocities and dimensions used for blockage corrections.
Adapted from Ref. [10].

Journal of Fluids Engineering

SEPTEMBER 2016, Vol. 138 / 091103-11

Downloaded From: http://asmedigitalcollection.asme.org/ on 08/07/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

a2

a4 b4 a4
b4 2a4  1

(A10)

Considering an unconfined actuator disk. The drag on the turbine is obtained through a momentum budget on a control volume
including the stream tube

Assuming no internal losses, the power generated by the turbine


is P Da2U. Substituting the expression obtained for drag (Eq.
(A9)) and the velocity through the turbine (Eq. (A10)), we get


1
b4 a4 b24  a 24
3
P q At U a4
2
b4 2a4  1

 2
1
CD
1
a4

a02


1
1 a04
2

(A21)

As seen from Eq. (A21), in the case of the unconfined actuator


disk, the flow through the disk is equal to the average of the
upstream and downstream velocities. Isolating a04 in Eq. (A21)
and using this expression in Eq. (A18), we get
(A22)



1 0
C a02 1  a02
4 D

Once the value of b4/a4 is converged, we compute a2 from the


results of Eqs. (A13) and (A12)

(A15)

In order to obtain the blockage correction, we want to determine the equivalent unconfined upstream velocity U 0 that would
lead (in the unconfined case) to a same power extracted than in
the confined case [1]. This implies a same flow velocity through
the turbine and a same drag on the turbine than those of the confined case.
So we basically have the following relations:
CD U2 C0D U 02

(A16)

a2 U a02 U0

(A17)

(A23)

Considering the hypothesis of Eqs. (A16) and (A17), Eq. (A23)


can be expressed as
U
a2

U 0 a 22 CD =4

(A14)

Fix an initial value for b4/a4 (for example, b4/a4 1.5).


Calculate a2/a4 from Eq. (A13).
Calculate 1/a4 from Eq. (A12).
Calculate b4/a4 from Eq. (A14).
Compute error between the new value of b4/a4 and the previous value.
(6) Repeat from step 2 until error on b4/a4 is negligible (i.e.,
less than 1  106).

091103-12 / Vol. 138, SEPTEMBER 2016

(A20)

Equaling those two equations for power we get

(A13)

(1)
(2)
(3)
(4)
(5)




1
a2 b4
a2 b4
e
1
a4 a4
a4 a4



1
q At U 03 a02 1  a04 2
2

Involving the drag coefficient, we obtain the following identity:

The blockage correction is obtained by an iterative process

a2

D 2 q At U 02 a02 1  a02

Finally, we express Eq. (A9) as a coefficient and isolating b4/a4


b4

a4

(A19)

and a second relation for power is obtained considering the


change in kinetic energy flux

(A12)

and then by expressing Eq. (A10) in terms of a4, we get a quadratic equation for a2/a4 which the positive solution (only the positive root has physical meaning) is
v
" 
#
u
2
u
b4
t
1 1 e
1
a4
a2



b
a4
e 41
a4

P q At U 03 a02 2 1  a04

(A11)

The blockage correction proposed by Bahaj et al. [4] uses the


same principles as the one developed above. Expressing continuity in terms of ratios involving the wake velocity deficit a4, isolating 1/a4, and introducing the blockage ratio e  At/A, we get


b4
1
a4

(A18)

A first relation for power is obtained from P D U 0 a02

Blockage Correction From Barnsley and Wellicome

1
b
a2
4e
a4 a4
a4

D q At U 02 a02 1  a04

(A24)

where a2 is known from Eq. (A15).


Corrected force, power, and operating conditions for the hypothetical unconfined case (noted with a 0 sign) are obtained from
the confined results using the following blockage corrections
involving the velocity ratio of Eq. (A24)
CP0 CP

 3
U
;
U0

CD0 CD

 2
U
;
U0

f0 f

 
U
U0

(A25)

References
[1] Glauert, H., 1947, The Elements of Aerofoil and Airscrew Theory, 2nd ed.,
Cambridge University Press, New York.
[2] Maskell, E. C., 1963, A Theory of Blockage Effects on Bluff Bodies and
Stalled Wings in Closed Wind Tunnel, Aeronautical Research Council, London, ARC R&M No. 3400.
[3] Pope, A., and Harper, J., 1966, Low-Speed Wind Tunnel Testing, 2nd ed.,
Wiley, New York.
[4] Bahaj, A., Molland, A., Chaplin, J., and Batten, W., 2007, Power and Thrust
Measurement of Marine Current Turbines Under Various Hydrodynamic Flow
Conditions in a Cavitation Tunnel and Towing Tank, Renewable Energy,
32(3), pp. 407426.
[5] Barnsley, M. J., and Wellicome, J. F., 1990, Final Report on the 2nd Phase of
Development and Testing of a Horizontal Axis Wind Turbine Test Rig for the
Investigation of Stall Regulation Aerodynamics, Carried Out Under ETSU
Agreement No. E.5A/CON5103/1746.
[6] Srensen, J. N., Shen, W. Z., and Mikkelsen, R., 2006, Wall Correction Model
for Wind Tunnels With Open Test Section, AIAA J., 44(8), pp. 18901894.
[7] Betz, A., 1920, Das Maximum der Theoretisch M
oglichen Ausn
utzung des
Windes Durch Windmotoren, Z. Gesamte Turbinenwesen, 26, pp. 307309.
[8] Lanchester, F. W., 1915, A Contribution to the Theory of Propulsion and the
Screw Propeller, Trans. Inst. Naval Archit., LVII, pp. 98116.
[9] Garret, C., and Cummins, P., 2007, The Efficiency of a Turbine in a Tidal
Channel, J. Fluid Mech., 588, pp. 243251.

Transactions of the ASME

Downloaded From: http://asmedigitalcollection.asme.org/ on 08/07/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

[10] Houlsby, G., Draper, S., and Oldfield, M., 2008, Application of Linear
Momentum Actuator Disc Theory to Open Channel Flow, Department of Engineering Science, University of Oxford, Report No. OUEL 2296/08.
[11] Whelan, J., Graham, J., and Peir
o, J., 2009, A Free-Surface and Blockage Correction for Tidal Turbines, J. Fluid Mech., 624, pp. 281291.
[12] Gauthier, E., Kinsey, T., and Dumas, G., 2013, RANS Versus Scale-Adaptive
Turbulence Modeling for Engineering Prediction of Oscillating-Foils
Turbines, 21th Annual Conference of the CFD Society of Canada, Sherbrooke,
Canada, May 69, p. CFDSC2013186.
[13] Kinsey, T., and Dumas, G., 2008, Parametric Study of an Oscillating Airfoil in
a Power Extraction Regime, AIAA J., 46(6), pp. 13181330.
[14] Kinsey, T., and Dumas, G., 2012, Computational Fluid Dynamics Analysis of
a Hydrokinetic Turbine Based on Oscillating Hydrofoils, ASME J. Fluids
Eng., 134(2), p. 021104.
[15] Kinsey, T., and Dumas, G., 2012, Optimal Tandem Configuration for
Oscillating-Foils Hydrokinetic Turbine, ASME J. Fluids Eng., 134(7),
p. 031103.
[16] Kinsey, T., and Dumas, G., 2012, Three-Dimensional Effects on an OscillatingFoils Hydrokinetic Turbine, ASME J. Fluids Eng., 134(7), p. 071105.

Journal of Fluids Engineering

[17] Kinsey, T., and Dumas, G., 2014, Optimal Operating Parameters for an Oscillating
Foil Turbine at Reynolds Number 500 000, AIAA J., 52(9), pp. 18851895.
[18] Zhu, Q., and Peng, Z., 2009, Mode Coupling and Flow Energy Harvesting by
a Flapping Foil, Phys. Fluids, 21(3), p. 033601.
[19] Zhu, Q., 2011, Optimal Frequency for Flow Energy Harvesting of a Flapping
Foil, J. Fluid Mech., 675, pp. 495517.
[20] Ashraf, M., Young, J., Lai, J., and Platzer, M., 2011, Numerical Analysis of an
Oscillating-Wing Wind and Hydropower Generator, AIAA J., 49(7),
pp. 13741386.
[21] CD-Adapco, 2014, STAR-CCM V9 User Guide, http://www.cd-adapco.com/
products/star-ccm
[22] Menter, F., 1994, Two-Equation Eddy-Viscosity Turbulence Models for Engineering Applications, AIAA J., 32(8), pp. 15981605.
[23] Bullivant, W. K., 1941, Tests of a NACA 0025 and 0035 Airfoils in the FullScale Wind Tunnel, Langley Memorial Aeronautical Laboratory, VA, NACA
Report No. 708.
[24] Kinsey, T., Dumas, G., Lalande, G., Ruel, J., Mehut, A., Viarouge, P., Lemay, J.,
and Jean, Y., 2011, Prototype Testing of a Hydrokinetic Turbine Based on Oscillating Hydrofoils, Renewable Energy, 36(6), pp. 17101718.

SEPTEMBER 2016, Vol. 138 / 091103-13

Downloaded From: http://asmedigitalcollection.asme.org/ on 08/07/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Das könnte Ihnen auch gefallen