Sunteți pe pagina 1din 10

Marco Raciti Castelli

e-mail: marco.raciticastelli@unipd.it
Ernesto Benini
e-mail: ernesto.benini@unipd.it
Department of Mechanical Engineering,
University of Padova, Via Venezia,
1-35131 Padova, Italy
Effect of Blade Inclination Angle
on a Darrieus Wind Turbine
This paper presents a model for the evaluation of energy performance and aerodynamic
forces acting on a small helical Darrieus vertical axis wind turbine depending on blade
inclination angle. It consists of an analytical code coupled to a solid modeling software
capable of generating the desired blade geometry depending on the desired design geo-
metric parameters, which is linked to a nite volume CFD code for the calculation of
rotor performance. After describing and validating the model with experimental data, the
results of numerical simulations are proposed on the bases of ve machine architectures,
which are characterized by an inclination of the blades with respect to the horizontal
plane in order to generate a phase shift angle between lower and upper blade sections of
0 deg, 30 deg, 60 deg, 90 deg, and 120 deg for a rotor having an aspect ratio of 1.5. The
effects of blade inclination on tangential and axial forces are rst discussed and then the
overall rotor torque is considered as a function of azimuthal position of the blades.
Finally, the downstream tip recirculation zone due to the nite blade extension is ana-
lyzed for each blade inclination angle, achieving a numerical quantication of the inu-
ence of induced drag on rotor performance, as a function of both blade element longi-
tudinal and azimuthal positions of the blade itself. DOI: 10.1115/1.4003212
1 Introduction and Background
Recent instabilities of world economy, due to the increasing
price of carbon-derivative fuels along with connected sociopoliti-
cal turbulences, have aroused the interest in the production of
renewable energy among the most industrialized western nations.
In this scenario, the continuous quest for clean energy is now
focusing on the local production of electric power, spread in a
wide area, so as to cooperate with the big electric power plants
located in just few specic strategic locations of the countries.
One of the most promising resources is wind power associated
with local production of clean electric power inside built environ-
ment such as industrial and residential areas, which has lead to the
development of the so called computational wind engineering.
This new discipline has also renewed the interest in vertical axis
wind turbines VAWTs, which present several advantages if com-
pared with the classical horizontal-axis wind turbines HAWTs,
primarily
lower cost;
lower need of maintenance;
lower sound emission;
independence from wind directions due to rotor axial-
symmetry;
better impact on the environment due to their
tridimensionality.
The vertical axis wind turbine has an inherently nonstationary
aerodynamic behavior mainly due to the continuous variation of
the blade angle of attack during the rotation of the machine: This
peculiarity involves the continuous variation both of the relative
velocity with respect to the blade prole andalthough to a lesser
extentof the corresponding Reynolds number. This phenom-
enon, typical of slow rotating machines, has a signicant effect
both on the dynamic loads acting on the rotor and on the gener-
ated power and, therefore, on performance.
Among others, Templin 1 rst developed a blade element-
momentum single streamtube numerical model to predict the per-
formance of a VAWT. Strickland 2,3 developed a blade element-
momentum multiple streamtube numerical model to predict the
performance of a VAWT rotor. Paraschivoiu 47 developed ana-
lytical and numerical aerodynamic models to investigate the per-
formance of VAWT, focusing on the phenomenon of dynamic
stall. Mertens 8 developed a blade element-momentum multiple
streamtube model to predict the performances of a fast rotating
VAWT in the skewed ow on a roof.
The complexity of the phenomena involved in the inherently
unsteady behavior of vertical axis turbines is often impossible to
investigate through classical aerodynamic tools, such as the theory
of the blade elements, and gives an account of the use of compu-
tational uid dynamics CFD aimed at determining the structure
of the ow eld vortices, three-dimensional effects, inuence of
spoke shape otherwise impossible to analyze, thanks to its inher-
ent ability to determine the aerodynamic components of actions
through the integration of the Navier-Stokes equations in the
neighborhood of the wind turbine.
Until now wind tunnel tests, involving considerable time and
nancial resources, have been the only way to fully characterize
the behavior of a rotor in order to obtain the operating torque
curves for the implementation of the control system.
In CFD simulations, the computer essentially replaces the
physical simulation in the wind tunnel, at least in principle. CFD
methods involve very large amounts of computation even for rela-
tively simple problems and their accuracy is often difcult to as-
sess when applied to a new problem where prior experimental
validation has not been done 9.
Performing CFD calculations provide knowledge about the
ow in all its details, such as velocities, pressure, temperature, etc.
Further, all types of useful graphical presentations, such as ow
lines, contour lines, and isolines, are readily available. This stage
can be compared with having completed a wind tunnel study or an
elaborate full-scale measurement campaign 10.
Ferreira et al. 11,12 investigated numerically the effect of dy-
namic stall in a 2D single-bladed VAWT, reporting the inuence
of the turbulence model in the simulation of the vortical structures
spread from the blade.
In this work a numerical methodology was developed in order
to predict the performance of a Darrieus rotor model as a function
of the phase shift angle between lower and upper blade sections.
Contributed by the International Gas Turbine Institute of ASME for publication in
the JOURNAL OF TURBOMACHINERY. Manuscript received September 3, 2010; nal
manuscript received September 4, 2010; published online July 15, 2011. Editor:
David Wisler.
Journal of Turbomachinery MAY 2012, Vol. 134 / 031016-1 Copyright 2012 by ASME
Downloaded From: http://asmedigitalcollection.asme.org/ on 09/30/2014 Terms of Use: http://asme.org/terms
2 Model Geometry
The aim of this work was to analyze numerically ve helical
Darrieus wind turbines, characterized by a different inclination of
the blades with respect to the horizontal plane.
All the helical blades analyzed lye on the surface of a 515 mm
radius cylinder, as shown in Fig. 1.
For the sake of clarity, the models were named according to the
phase shift angle between lower and upper blade sections in the
case of an equivalent rotor having aspect ratio H/ 2R =1.5. The
numerically tested rotor architectures were instead characterized
by an aspect ratio equal to 1 in order to reduce the total number of
mesh elements. Table 1 shows the name and description of the
ve models, together with their blade inclination with respect to
the horizontal plane.
A single-bladed rotor was analyzed, mainly for two reasons
to reduce the total number of mesh elements and thereafter
the computational time;
to highlight the behavior of a single-blade without the dis-
torting effects due to the presence of other blades.
The common features of the tested rotors are summarized in
Table 2, both for the real model characterized by an aspect ratio
of 1.5 and for the computational one characterized by an aspect
ratio of 1.
Figure 2 shows a comparison between Model 0, Model 60, and
Model 120 blades.
The blade azimuthal position is identied by the angular coor-
dinate of the pressure center of the blade midsection, as can be
seen in Fig. 3. For Model 0 conguration, this coordinate also
identies the position of all blade sections.
Figure 4 shows a schematic of the survey methodology utilized,
consisting in the coupling of an analytical code to a solid model-
ing software, capable of generating the desired blade geometry
depending on the desired design geometric parameters, which is
linked to a nite volume CFD code for the calculation of rotor
performance.
3 Computational Model and Validation
Before analyzing the models described in the previous section,
a complete validation work based on wind tunnel measurements
has been conducted 13. The experimental setup consisted in a
classical vertical-bladed Darrieus rotor made of aluminum and
carbon bers using a NACA 0021 blade prole with a chord
length of 85.8 mm, which was tested in Bovisas low turbulence
facility Milan.
A computational domain of rectangular shape has been chosen,
having the same wind tunnel external size, as can be seen in Fig.
5: The wall boundary conditions of the model consisted in two
lateral walls spaced 2000 mm apart from the wind tunnel center-
Fig. 1 Exemplication of a helical blade developing on the
surface of a cylinder
Table 1 Main features of the ve models
Model name

deg

deg
0 0 90.00
30 20 80.10
60 40 70.76
90 60 62.36
120 80 55.08
Table 2 Comparison between real model and computational
model
Real model Computational model
Horizontal section NACA 0021 NACA 0021
Chord, c 85.8 mm 85.8 mm
Rotor radius, R 515 mm 515 mm
Rotor height, H 1545 mm 1030 mm
Fig. 2 Comparison between Model 0, Model 60, and Model 120
blade
Fig. 3 Azimuthal coordinate
031016-2 / Vol. 134, MAY 2012 Transactions of the ASME
Downloaded From: http://asmedigitalcollection.asme.org/ on 09/30/2014 Terms of Use: http://asme.org/terms
line the wind tunnel measured 4000 mm in width and 3880 mm
in height. The rotor axis was placed on the symmetry position of
the wind tunnel section. 3D simulations were performed in order
to take into account also the drag effect induced by the spokes.
Only half of the experimental setup was modeled due to its verti-
cal symmetry: In this case a symmetry boundary condition was
used. Anyway, the geometrical features of the model did not allow
other simplications to be performed. Table 3 shows the valida-
tion model main features. The effect of gravity on the rotor work-
ing curves has not been contemplated, being considered not inu-
ential for the scope of this work.
Inlet and outlet boundary conditions were placed respectively
10 diameters upwind and 14 diameters downwind of the rotor,
allowing a full development of the wake, as suggested by the
work of Ferreira et al. 11.
The correction due to wind tunnel blockage was not applied in
order to minimize any sources of error due to a wrong estimation
of the blockage of the wind tunnel itself. Furthermore, this choice
has the signicant advantage of reducing the computational do-
main, allowing a saving in the total number of mesh elements. The
correction of the friction resistive torque due to the bearings was
taken into account.
As the aim of the present work was to reproduce the operation
of a rotating machine, the use of moving sub-grids was necessary.
The simulation domain was divided in two sub-grids
Rotor sub-grid, rotating with angular velocity Fig. 6;
Wind Tunnel sub-grid, xed.
An isotropic unstructured mesh was chosen for the Rotor sub-
grid in order to guarantee the same accuracy in the prediction of
rotors performance during the rotationaccording to the studies
of Commings, Forsythe, Morton, and Squires 14and also in
order to test the prediction capability of a very simple grid. Con-
sidering their features of exibility and adaption capability, un-
structured meshes are in fact very easy to obtain, also for complex
geometries, and often represent the rst attempt in order to get
a quick response from the CFD in engineering work.
An unstructured mesh was chosen also for the Wind Tunnel
sub-grid in order to reduce engineering time to prepare the CFD
simulations. An interface boundary condition was assigned to the
interface between the Rotor sub-grid and the Wind Tunnel sub-
grid.
In order to control the size of mesh elements near the surface of
the blade, the latter was placed inside a 400 mm diameter control
cylinder Fig. 7. An interior boundary condition was assigned to
the interface between the control cylinder and the remaining Rotor
sub-grid mesh.
The computational grids were constructed from lower topolo-
gies to higher ones, adopting appropriate size functions Fig. 8, in
order to cluster grid points near the leading edge and the trailing
edge of the blade prole, so as to improve the CFD code capabil-
ity of determining lift, drag, and the separation of the ow from
the blades itself. Mesh density was also based on the local curva-
ture of the blade elements.
As a nal step, the mesh elements have been fully converted
into polyhedra. This option, applicable to unstructured mesh of
tetrahedral type, has the advantage of reducing the total number of
grid elements, producing in the same time greater mesh regularity,
as shown in Figs. 9 and 10.
Because conversion into polyhedra is a very resource intensive
process, two separate les were created for the Rotor sub-grid and
the Wind Tunnel sub-grid. It was thus possible to convert the two
Fig. 4 Schematic of the survey methodology
Fig. 5 Computational domain validation model
Table 3 Validation model main features
Prole type NACA 0021
c 85.8 mm
R 515 mm
H 1456.4 mm
A 1.236 m
2
0.25
Spoke-blade connection 0.5 c
Wind tunnel dimensions 40003800 mm
2
Fig. 6 Rotor sub-grid mesh validation model
Fig. 7 Control cylinder validation model
Journal of Turbomachinery MAY 2012, Vol. 134 / 031016-3
Downloaded From: http://asmedigitalcollection.asme.org/ on 09/30/2014 Terms of Use: http://asme.org/terms
main grid areas independently and then to reassemble them into a
single le.
The conversion into polyhedra has achieved a 6070% reduc-
tion in the total element number. On the other hand, a polyhedra
grid occupies the memory of a tetraheda almost twice. The total
budget is therefore favorable to polyhedra, even considering the
fact that a polyhedral mesh shows more marked regularity fea-
tures than the corresponding tetrahedral one and therefore allows
a much faster convergence.
In order to test the code sensitivity to the number of grid points,
three unstructured meshes were adopted for the Rotor sub-grid,
while the Wind Tunnel sub-grid remained substantially the same.
After some corrections to take into account spoke drag, the
average torque values measured in the wind tunnel for a 9 m/s
wind speed and different tip speed ratios were compared with
those obtained from CFD analysis for the three different grids
characterized by different blade size function values and three
different turbulence models k- SST, k- realizable, and Spalart
Allmaras Table 4.
Figures 11 and 12 show the evolution of the instantaneous
torque coefcient, dened as
C
T
=
M
1/2 A V

2
R
1
as a function of the azimuthal position for two different meshes
and for the three adopted turbulence models.
The nal choice was mesh mod A, based on a better distribution
of the blade y
+
parameter and k- SST turbulence model because
of its better ability to describe ow separation 1517, which
occurs in ow elds dominated by adverse pressure gradients,
even if it has shown a certain sensitivity to grid size 18. Stan-
dard wall functions were used to model the boundary layer.
The temporal discretization has been achieved by imposing a
time step equal to the lapse of time the rotor takes to make a 1 deg
rotation. An improved spatial-discretization simulation did not
show any signicant variation.
The commercial CFD package used was FLUENT 6.3.26, which
implements 3-D Reynolds-averaged NavierStokes equations us-
ing a nite volume-nite element based solver. The uid was as-
Fig. 8 Blade size functions
Fig. 9 Improved mesh regularity after conversion into polyhe-
dra 1
Fig. 10 Improved mesh regularity after conversion into poly-
hedra 2
Table 4 Grid and turbulence models used for calculations
Mesh
Maximum blade element size
mm Turbulence model
Mod 0 2 k- k- SST S-A
Mod A 3 k- k- SST S-A
Mod B 3.5 k- k- SST S-A
Fig. 11 Effect of grid resolution on the instantaneous torque
for a single-bladed rotor turbulence model: k- SST
Fig. 12 Effect of turbulence model on the instantaneous
torque for a three-bladed rotor mod A mesh
031016-4 / Vol. 134, MAY 2012 Transactions of the ASME
Downloaded From: http://asmedigitalcollection.asme.org/ on 09/30/2014 Terms of Use: http://asme.org/terms
sumed to be incompressible, being the maximum uid velocity on
the order of 60 m/s.
The simulations, performed on an 8 processor, 2.33 GHz clock
frequency computer, have been run until the instantaneous torque
values showed a deviation of less than 1% compared with the
corresponding values of the previous period. Total CPU time has
been about 20 days for each simulation.
4 Model Analysis
The ve helical models have kept some common points with
the validation model, particularly as far as the chord length and
the rotor radius are concerned. The blade-spoke connection point
has been changed, since it had been placed in the center of pres-
sure of the prole, corresponding to 25% of the chord length
behind the blade leading edge, while in the validation model it had
been placed close to 50%. The analysis of the impact of this
variation on the parameter y
+
showed no signicant effect, mainly
due to the fact that in fast rotating machines, the values of blade
relative velocity are determined primarily by the angular velocity
of the rotor itself and to a lesser extent by the speed of undis-
turbed air ow.
In order to control the size of mesh elements near the surface of
the blade, the latter was placed inside a control cylindroid, which
was in turn subdivided into 20 equal subvolumes, numbered from
1 to 20 from top downward Fig. 13. This choice has also proved
extremely useful in order to study blade tip effects, allowing to
analyze the contributions to torque generation of any single blade
subelement.
The validation model was made of straight blades: It was there-
fore possible to proceed to the analysis of only half the model,
exploiting the symmetry with respect to the median horizontal
plane. The helical models do not show any symmetry: It was
therefore necessary to simulate the whole rotor.
The external size of the computational domain was changed: In
the validation model the Wind Tunnel sub-grid had to reproduce
the geometry of the wind tunnel Table 5. In order to avoid block-
age effects due to the proximity to wind tunnel walls, the compu-
tational domain for the ve helical models has been enlarged,
allowing to analyze the behavior of the rotor in an open ow eld.
This additional change from validation model was limited to the
outer portions of the computational domain and it was therefore
considered negligible with respect to the values of blade y
+
Figs.
14 and 15.
The decision to increase the size of the computational domain
has led to an unusual hourglass shape for the Rotor sub-grid. In
order to avoid a rotating cylindrical grid as tall as the whole com-
putational domain, a short cylindrical grid was connected to the
upper and lower surfaces using two truncated cone elements. The
mesh element inside these cones are characterized by the same
growth rate from the lateral surface of the central cylinder to the
outer surface of the computational domain.
Figure 16 shows the central element of the Rotor sub-grid. The
clustering of grid points inside the Rotor sub-grid can be seen.
The choice to extend the rotating mesh to the upper and lower
surfaces of the computational domain was dictated by the need not
Fig. 13 Blade subdivision into 20 zones, numbered from 1 to
20 from top downward
Table 5 Comparison between validation model and computa-
tional model
Validation model Computational model
c 85.8 mm 85.8 mm
R 515 mm 515 mm
H 1456.4 mm 1030 mm
Spoke-blade connection 0.5 c 0.25 c
Fig. 14 Computational domain and relative mesh computa-
tional model, 1
Fig. 15 Computational domain and relative mesh computa-
tional model, 2
Fig. 16 Rotor sub-grid mesh for Model 0 computational
model
Journal of Turbomachinery MAY 2012, Vol. 134 / 031016-5
Downloaded From: http://asmedigitalcollection.asme.org/ on 09/30/2014 Terms of Use: http://asme.org/terms
to create interfaces parallel to the principal ow direction, which
could cause numerical problems especially with regard to Omega
residuals, as can be seen in Figs. 17 and 18.
5 Results and Discussion
Figure 19 represents the ve helical model power curves for an
incident wind speed of 9 m/s as a function of the tip speed ratio,
dened as
= R/V

2
The power coefcients do not show signicant differences
since the phase shift angle between lower and upper blade sec-
tions is lower than 60 deg. For Model 90 and Model 120 the
power coefcients show a marked decrease and the optimum tip
speed ratios show a small increase.
Figure 20 shows the distribution of instantaneous torque coef-
cient as a function of azimuthal position of rotor Model 0, Model
60, and Model 120, for a tip speed ratio of 3.36. Once more,
signicant differences depending on the phase shift angle between
lower and upper blade sections are visible.
The rst maximum torque peak at 96 deg azimuthal position
is however the same for the three models, while the second peak
close to 276 deg azimuthal position is slightly spaced backward
for Model 60 and Model 120.
Table 6 shows the values of instantaneous torque coefcient for
the two corresponding azimuthal positions 92 deg and 276 deg.
The peak value decrease is very low about 5% between Model 0
and Model 60 but increases up to 24% between Model 0 and
Model 120.
Figures 2123 compare the contributions of instantaneous
torque in each of the 20 blade zones for the two described above
peak torque azimuthal positions and for a third intermediate azi-
muthal position 48 deg.
The Model 0 straight blade exhibits a symmetrical torque pro-
Fig. 17 Numerical problems caused by parallel to principal
ow direction interfaces and problem solution using extended
rotating mesh 1
Fig. 18 Numerical problems caused by parallel to principal
ow direction interfaces and problem solution using extended
rotating mesh 2
Fig. 19 Power curves for the ve models
Fig. 20 Instantaneous torque coefcient as a function of azi-
muthal coordinate
Table 6 Instantaneous torque coefcients for azimuthal posi-
tions of 92 deg and 276 deg Model 0, Model 60, and Model 120
Azimuthal position deg 92 276
Model 0 C
T
0.201 0.049
Model 60 C
T
0.191 5% 0.046
Model 120 C
T
0.152 24.4% 0.042
Fig. 21 Contribution of instantaneous torque in each blade
zone for Model 0 and Model 120 azimuthal coordinate 92 deg
031016-6 / Vol. 134, MAY 2012 Transactions of the ASME
Downloaded From: http://asmedigitalcollection.asme.org/ on 09/30/2014 Terms of Use: http://asme.org/terms
le with respect to the blade centerline a decrease in blade per-
formance can also be registered due to vortex shedding in corre-
spondence to the upper and lower blade tips. This is not the case
for helical blades as a consequence of the choice for the reference
system in dening the blade azimuthal position: In a straight
blade, each zone features the same azimuthal position with respect
to the relative ow eld, while in the case of helical blades it has
a different angular position with respect to the blade centerline.
Two torque peaks are visible in Fig. 22 for Model 0, located
close to zone 3 and zone 18 because Model 00 is symmetrical,
the torque distribution is also symmetrical: The cause of this
phenomenon is not yet clear; however, we suggest that for some
azimuthal locations, a vertical air suction into the rotor occurs,
which accelerates the ow eld in the proximity of blade tips.
A more appropriate representation of the torque generated by
each blade zone can be ascertained in Fig. 24, where the average
value of the torque coefcient generated during a complete 360
deg rotation is represented as a function of the blade zone itself:
Most of the torque in a helical blade is produced from the bottom
blade zones, while the upper blade zone contribution is negative.
The cause of this phenomenon is not easy to understand; how-
ever, it can be argued that tip effects play a dominant role at the
top than at the bottom of the blade. This aspect is analyzed in the
following.
Figure 25 shows the instantaneous torque values as a function
of azimuthal position for zone 1 and zone 20 for Model 0 and
Model 120.
As can be clearly seen, the torque generated by the lower blade
zone is much higher than the torque generated by the correspond-
ing upper blade zone Fig. 26. However, the velocity-vector vi-
sualizations of Figs. 27 and 28 for upper and lower blade tip zones
show no appreciable differences in the vortex shedding.
The explanation of the phenomenon lies therefore in other fac-
tors. Figures 29 and 30 visualize the pathlines for Model 0 and
Model 60 in correspondence to the blade centerline zone 10,
which is far from the perturbed areas due to tip effects. Model 0
streamlines are parallel to the horizontal plane, while Model 60
streamlines clearly deviate upward.
The above description shows that, close to the blade, the ow
eld takes a direction perpendicular to the leading edge, according
to Fig. 31.
The reduced contribution to torque generation of the upper
blade element is caused by the fact that pathlines are unable to
Fig. 22 Contribution of instantaneous torque in each blade
zone for Model 0 and Model 120 azimuthal coordinate 276 deg
Fig. 23 Contribution of instantaneous torque in each blade
zone for Model 0 and Model 120 azimuthal coordinate 48 deg
Fig. 24 Average torque for each blade zone for Model 0 and
Model 1200
Fig. 25 Instantaneous torque coefcient values as a function
of azimuthal position for Model 0 and Model 120 zone 1
Fig. 26 Instantaneous torque coefcient values as a function
of azimuthal position for Model 00 and Model 120 zone 20
Journal of Turbomachinery MAY 2012, Vol. 134 / 031016-7
Downloaded From: http://asmedigitalcollection.asme.org/ on 09/30/2014 Terms of Use: http://asme.org/terms
fully develop along a blade prole because of an abrupt halt due
to the horizontal cut in the blade itself, as can be seen in Fig. 32.
Even the lower blade element presents a similar cut, but since
the pathlines bend upward, it does not affect the continuity of ow
along the blade element itself: Therefore, no reduction in the
torque is seen.
The deviation of streamlines is also responsible for the marked
decrease in rotor performance for high values of phase shift angle
between lower and upper blade sections. Figure 33 shows the
distortion of the blade section in direction perpendicular to the
leading edge due to blade curvature, compared with the original
NACA 0021 blade section in the horizontal plane.
For high phase shift angle, the ow eld is no more interacting
with a NACA prole but with a deformed one, with consequent
decrease of overall rotor efciency.
Blade deformation can also be seen from Table 7, comparing
the chord length of the blade section interacting with the ow eld
with the original NACA 0021 section for the ve analyzed models
Fig. 34.
The reduction of chord length, and the consequent reduction of
rotor solidity, are also responsible for the slight increment in op-
timum tip speed ratio values with the phase shift angle between
lower and upper blade sections, as described in Fig. 19.
Figure 35 shows the axial thrust coefcient acting on blades for
Model 0 and Model 120, dened as
C
Fz
=
F
z

1/2 A V

2
3
as a function of the azimuthal position for a tip speed ratio of
3.36. While there is no straight blade axial thrust, the helical
blades produce a downward axial thrust increasing the axial bear-
ing load.
Fig. 27 Velocity vectors visualization for upper and lower
blade tip zones 1
Fig. 28 Velocity vectors visualization for upper and lower
blade tip zones 2
Fig. 29 Model 00, zone 10: streamlines are parallel to the hori-
zontal plane
Fig. 30 Model 0, zone 10: streamlines deviate upward
Fig. 31 Streamlines deviation in a direction perpendicular to
the leading edge
Fig. 32 Horizontal cut on the top of helical blades
031016-8 / Vol. 134, MAY 2012 Transactions of the ASME
Downloaded From: http://asmedigitalcollection.asme.org/ on 09/30/2014 Terms of Use: http://asme.org/terms
The explanation for this phenomenon lies once more in the fact
that pathlines bend upwards in a direction orthogonal to the lead-
ing edge. This causes, for Newtons laws of motion, a uid down-
ward thrust on the blade.
In order to reduce axial bearing loads, the helical blade should
develop so that the lower sections meet the ow before the upper
ones.
6 Conclusions and Future Work
In this paper, a model for the evaluation of energy performance
and aerodynamic forces acting on a small helical Darrieus vertical
axis wind turbine depending on blade inclination angle has been
developed, based on an analytical code coupled to a solid model-
ing software that was linked to a nite volume CFD code for the
calculation of rotor performance.
The obtained results, based on ve machine architectures,
which are characterized by an inclination of the blades with re-
spect to the horizontal plane in order to generate a phase shift
angle between lower and upper blade sections of 0 deg, 30 deg, 60
deg, 90 deg, and 120 deg for a rotor with aspect ratio of 1.5,
demonstrate that the average torque does not change substantially
for lower phase shift angle but shows a marked decrease for high
values of phase shift angle between upper and lower blade sec-
tions.
Up to 60 deg phase shift angle, torque peaks showed a slight
decrease, depending on blade inclination, but no signicant reduc-
tion of pulsating torque depending on the phase shift angle be-
tween lower and upper blade sections has been proved.
The obtained results demonstrate also that, close to the blade,
the ow eld takes a direction perpendicular to the leading edge,
with streamlines clearly deviating upward for a helical blade de-
veloping so that the upper sections meet the ow before the lower
ones.
The obtained results demonstrate also that a signicant contri-
bution to the mean torque in a helical blade is produced from the
bottom blade zones. Although the reason for this phenomenon is
not easy to understand, it has been assumed to be caused by the
fact that pathlines are unable to fully develop along a upper blade
prole because of an abrupt halt due to the horizontal cut in the
blade itself.
Finally, it has been demonstrated that helical blades developed
so that the upper sections meet the ow before the lower ones
produce a downward axial thrust increasing the axial bearing load.
The explanation for this phenomenon lies once more in the fact
that pathlines bend upward in a direction orthogonal to the leading
edge. This causes, for Newtons laws of motion, a uid downward
thrust on the blade.
Some aspects still remain to be investigated: First of all, nu-
merical analysis need to be extended in order to examine the
effect of constructing the blade prole using inclined normal to
leading edge airfoil sections, which could achieve better perfor-
mance, due to streamline inclination.
Also, the cause of the two torque peaks close to tip blades for
some azimuthal position and the corresponding possible vertical
air suction into the rotor remain to be further investigated.
Finally, an accurate analysis is needed to investigate axial loads
acting on a helical blade developed so that the lower sections meet
the ow before the upper ones, thus reducing the total loads acting
on the bearings.
Nomenclature
A rotor swept area, m
2
c blade chord, mm
C
Fz
axial thrust coefcient
C
p
rotor power coefcient
C
T
instantaneous torque coefcient
F rotor axial force, N
H rotor height, mm
P rotor power, W
Fig. 33 Blade section interacting with the ow eld dark blue
compared with the original NACA 0021 horizontal section light
blue for phase shift angle of 120 deg
Table 7 Comparison between chord lengths of the blade sec-
tion interacting with the ow eld for the ve models analyzed
Model name
c
normal to blade
mm Variation %
0 85.8 0.00
30 84.5 1.51
60 81.0 5.59
90 76.0 11.42
120 70.3 18.06
Fig. 34 Blade section distortion dark blue for phase shift
angle of 60 deg and 120 deg, compared with the original NACA
0021 section light blue
Fig. 35 Axial forces acting on Model 0 and Model 60 blades for
a tip speed ratio of 3.36
Journal of Turbomachinery MAY 2012, Vol. 134 / 031016-9
Downloaded From: http://asmedigitalcollection.asme.org/ on 09/30/2014 Terms of Use: http://asme.org/terms
R rotor radius, mm
T rotor torque, Nm
V

undisturbed ow velocity, m/s


phase shift angle between lower and upper
blade sections, deg
blade inclination with respect to the horizontal
plane, deg
tip speed ratio
blade azimuthal coordinate, deg
air density assumed 1.225 kg/ m
3
, kg/ m
3
rotor solidity
rotor angular velocity, rad/s
References
1 Templin, R. J., 1974, Aerodynamic Theory for the NRC Vertical-Axis Wind
Turbine, NRC of Canada, Report No. TR LTR-LA-160.
2 Strickland, J. H., The Darrieus Turbine: A Performance Prediction Model
Using Multiple Streamtube, SAND75-0431.
3 Oler, J. W., Strickland, J. H., Im, B. J., and Graham, G. H., Dynamic Stall
Regulation of the Darrieus Turbine, SAND83-7029.
4 Allet, A., and Paraschivoiu, I., 1995, Viscous Flow and Dynamic Stall Effects
on Vertical-Axis Wind Turbines, Int. J. Rotating Mach., 21, pp. 114.
5 Brahimi, M. T., Allet, A., and Paraschivoiu, I., 1995, Aerodynamic Analysis
Models for Vertical-Axis Wind Turbines, Int. J. Rotating Mach., 21, pp.
1521.
6 Masson, C., Leclerc, C., and Paraschivoiu, I., 1998, Appropriate Dynamic-
Stall Models for Performance Predictions of VAWT With NLF Blades, Int. J.
Rotating Mach., 42, pp. 129139.
7 Paraschivoiu, I., 2002, Wind Turbine Design: With Emphasis on Darrieus Con-
cept, Polytechnic International Press, Montreal.
8 Mertens, S., van Kuik, G., van Bussel, G., Performance of a High Tip Speed
Ratio H-Darrieus in the Skewed Flow on a Roof, Paper No. AIAA-2003-
0523.
9 Stathopoulos, T., 2004, Wind Effects on People, Proceedings of the Interna-
tional Conference on Urban Wind Engineering and Building Aerodynamics
Impact of Wind Storm on City Life and Built Environment, COST Action C14,
von Karman Institute, Rhode-Saint-Gense, Belgium.
10 Jensen, A. G., Franke, J., Hirsch, C., Schatzmann, M., Stathopoulos, T., Wisse,
J., and Wright, N. G., 2004, CFD TechniquesComputational Wind Engi-
neering, Proceedings of the International Conference on Urban Wind Engi-
neering and Building AerodynamicsImpact of Wind and Storm on City Life
and Built EnvironmentWorking Group 2, COST Action C14, von Karman
Institute, Rhode-Saint-Gense, Belgium.
11 Simao Ferreira, C. J., Bijl, H., van Bussel, G., and van Kuik, G., 2007, Simu-
latine Dynamic Stall in a 2D VAWT: Modeling Strategy, Verication and Vali-
dation With Particle Image Velocimetry Data, The Science of Making Torque
from Wind, Journal of Physics: Conference Series 75.
12 Simao Ferreira, C. J., van Bussel, G., Scarano, F., and van Kuik, G., 2007, 2D
PIV Visualization of Dynamic Stall on a Vertical Axis Wind Turbine, AIAA,
Reston, VA.
13 Raciti Castelli, M., Pavesi, G., Battisti, L., Benini, E., Ardizzon, G., 2010,
Modeling Strategy and Numerical Validation for a Darrieus Vertical Axis
Micro-Wind Turbine, ASME Paper No. IMECE2010-39548.
14 Cummings, R. M., Forsythe, J. R., Morton, S. A., and Squires, K. D., 2003,
Computational Challenges in High Angle of Attack Flow Prediction, Prog.
Aerosp. Sci., 395, pp. 369384.
15 Spalart, P. R., 1994, A One-Equation Turbulence Model for Aerodynamic
Flows, Rech. Aerosp., 1, pp. 521.
16 Menter, F. R., 1994, Two-Equation Eddy-Viscosity Turbulence Models for
Engineering Applications, AIAA J., 328, pp. 1598.
17 Wilcox, D. C., 1998, Turbulence Modeling for CFD, DCW Industries Inc., La
Canada, CA.
18 Bardina, J. E., Huang, P. G., and Coakley, T. J., 1997, Turbulence Modeling
Validation, Testing and Development, NASA, Technical Report 110446.
031016-10 / Vol. 134, MAY 2012 Transactions of the ASME
Downloaded From: http://asmedigitalcollection.asme.org/ on 09/30/2014 Terms of Use: http://asme.org/terms

S-ar putea să vă placă și