You are on page 1of 7

Takao Maeda

e-mail: maeda@mach.mie-u.ac.jp

Yasunari Kamada
e-mail: kamada@mach.mie-u.ac.jp

Jun Suzuki
e-mail: jun@fel.mach.mie-u.ac.jp

Hideyasu Fujioka
e-mail: fujioka@fel.mach.mie-u.ac.jp
Division of Mechanical Engineering,
Graduated School of Mie University,
1577 Kurimamachiya, Tsu,
Mie 514-8507, Japan

Rotor Blade Sectional


Performance Under Yawed Inflow
Conditions
This study shows the results of pressure distribution measurements on a rotor blade of a
horizontal axis wind turbine under various yawed operations. The experiments are carried out in a wind tunnel with a 2.4 m diameter test rotor. In the measurements, the power
curve and pressure distributions are measured for different azimuth angles. By increasing
yaw angle, the maximum value of power coefficient of the rotor decreases. The sign of the
yaw angle does not have any effect on power performance. The aerodynamic forces are
discussed using the axial and rotational force coefficients for each azimuth angle. In the
case of higher tip speed ratios, the blade section passing on the upstream side in yawed
operations has a greater contribution to the rotor torque than that on the downstream
side. In this tip speed range, the aerodynamic forces at the 70% radius section appear
proportional to the angle of attack. In the case of the lower tip speed ratios, the blade on
the downstream side does not contribute to rotor torque, which appears to result from
separation. DOI: 10.1115/1.2931514
Keywords: wind turbine, performance, pressure distribution, yawed inflow

Introduction

Aerodynamic efficiency of wind turbines greatly depends on


the performance of the rotor blade, the airfoil section, and the plan
form. There are many studies on wind turbine aerodynamics 1,2.
Performance of rotor blades is evaluated by velocity measurements of the flow around rotating blades and also by pressure
distribution measurements. The latter method can detect the forces
acting on the blade section directly. Therefore, pressure measurements play a particularly important role in grasping aerodynamic
performance of rotating blades. There are a number of studies
using pressure distribution measurements during rotation in field
conditions 3,4 and in the wind tunnel 5, but there are still many
unanswered questions. In the field, the inflow wind is not steady
in both direction and velocity. The wind turbine yaw mechanism
changes the nacelle direction in the wind direction every 10 min.
With fluctuating winds, wind turbines operate in some yaw misalignment most of the time. Therefore, it is important to understand the flow phenomenon at the rotor in yawed conditions.
There are many relevant numerical studies 6,7. Imamura et al.
6 investigated the local angle of attack to include the viscous
effects by the airfoil sectional data and shows rotor performance
in the yawed condition. Pesmajoglou et al. 7 modified the model
for the vortex lattice method to simulate the effect of stall. Detailed measurements for the yawed rotor in the wind tunnel are
reported in Refs. 810. Haans et al. 8 performed detailed measurements of the tip vortex location behind the yawed rotor D
= 1.2 m using quantitative flow visualization. Medici et al. 9
studied the velocity field in the wake of the rotor D = 0.18 m
using a two-component hot wire. Grant et al. 10 measured the
motion of the shed vorticity of the rotor D = 0.9 m by laser-sheet
flow visualization. In this study, we operated a 2.4 m diameter
wind turbine rotor in yawed conditions in a wind tunnel, and
investigated the aerodynamic performance of the rotor blade by
measuring the pressure distribution on the blades at specific radial
Contributed by the Solar Energy Engineering Division for publication in the JOURSOLAR ENERGY ENGINEERING. Manuscript received June 2, 2007; final manuscript received November 26, 2007; published online July 16, 2008. Review conducted by Spyros Voutsinas. Paper presented at the European Wind Energy
Conference and Exhibition 2007, Milan, Italy.

NAL OF

Journal of Solar Energy Engineering

positions. The full operational range of tests including low tip


speed ratios is presented in this paper. The variation of the aerodynamic force acting on the blade element under yaw at r / R
= 0.7 for each azimuth angle is clarified experimentally.

Experimental Apparatus and Methods

The experimental apparatus is shown in Fig. 1 11,12. A single


return type wind tunnel with an outlet diameter of 3.6 m was used
in this study. The maximum wind speed was 30 m / s. The uniformity of the flow velocity is less than 1.0% in the measurement
section. The turbulence intensity is less than 0.5%. The rotor diameter of model wind turbine is D = 2.4 m. The test wind turbine
is three bladed and of upwind horizontal axis type. Test blades
were designed similar to commercial wind turbine blades, featuring twist, taper, and four different airfoil sections along the span.
The test rotor was placed 1D downstream from the nozzle of the
wind tunnel. The test wind turbine nacelle had a gearbox allowing
to change rotational speed, a variable speed generator, a torque
meter, a rotational speed sensor, and a sensor for detecting the
azimuth angle. The pressure sensor array, which was accommodated in the rotor hub, is of semiconductor type, having 32 measuring ports and a reference-pressure port. To avoid the effect of
the centrifugal force on the sensor, the sensing membrane surface
was set parallel to the rotating plane. Figure 2 shows the locations
of the instrumented sections: 0.3R, 0.5R, 0.7R, and 0.9R, and the
arrangement of pressure taps over a section. The geometrical
specifications of the test blade are shown in Table 1. The 32 pressure taps were set on the blade surface for each measuring section
that is perpendicular to the pitch axis. The diameter of the tap was
0.4 mm.
The test wind speed was fixed at U0 = 7 m / s, while the pitch
angle was set at = 2 deg, which corresponds to optimum power
output for zero yaw. The blade pitch angle is the angle between
the rotor disk surface and blade tip chord, and is positive when the
blade leading edge is inclined upstream. Experiments were performed with seven yaw angle settings, = 0 , 15, 30, 45 with plus
and minus. The yaw angle is the angle between the rotor axis and
the wind direction. Figure 3 shows the direction of rotation, the
position of zero azimuth at 12 o clock, as well as a positive yaw
angle. The rotational speed of the rotor was changed from

Copyright 2008 by ASME

AUGUST 2008, Vol. 130 / 031018-1

Downloaded 06 Dec 2011 to 193.204.249.101. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Wind Tunnel nozzle

4500

Inlet

Azimuth Angle

Wind Turbine

2400

3600

Yaw Angle
Wind Turbine

Wind

1D

Fig. 3 Definition of azimuth and yaw angles

0.5

1.0

0.5 1.0

Fig. 1 Experimental setup

87 rpm to 352 rpm = 1.6 6.5 during the tests. Pressure data
were averaged 32 times at the same azimuth angle BIN. Azimuth
angles used in the discussion are = 0 deg, 90 deg, 180 deg, and
270 deg. The Reynolds number, which is based on the chord at the
blade tip and the tip speed, was about 2.1 105 at the optimum tip
speed ratio, = 5.20. The wind speed was measured by a Pitot
tube set upstream of the rotor in the tunnel test section. The position of the Pitot tube 1.5D upstream from the rotor was selected to represent average velocity over the test section without
the effect of the rotor induced velocity.

Fig. 2 Instrumentation of the rotor blade: a radial placement


of the instrumented sections; b position of pressure taps
r / R = 0.7

Aerodynamics Forces Acting on the Rotating Blade

Usually aerodynamic forces acting on a blade cross section are


evaluated by means of the lift force coefficient Cl and the drag
force coefficient Cd both depending on the inflow direction Fig.
4. However, a wind turbine blade is rotating and the relative
inflow direction for blade elements cannot be easily measured.
Therefore, in this study, the axial force coefficient Ca, which is
defined in the rotating axis direction, and the rotational force coefficient Cr, which is defined in the rotation direction, were used
as parameters to evaluate the aerodynamic force in rotating state.

Dynamic Characteristics of Pressure Scanner

Pressure measurements on the blade surface were performed by


means of a high-speed pressure scanner Scanivalve ZOC22B
fitted in the hub. The pressure taps are pneumatically connected to
the pressure scanner through the copper pipes and flexible tubes.
When the wind turbine was operated under yawed conditions, the
surface pressure changed dynamically with the azimuth angle.
Therefore, the pressure system was calibrated dynamically. In order to examine the effect of tubing length on the pressure signal,
the pressure signals of both a direct connection to the scanner and
through the tubing were compared. The amplitude of the oscillating pressure was 200 500 Pa, while the frequency changed up to
50 Hz. Figure 5 shows the relative amplitude and phase delay
through the pressure system compared to those obtained from the
direct connection to the pressure scanner. The amplitude decreases
as the frequency increases. The rotational frequency of the test
rotor is 5 Hz at the optimum tip speed ratio, and the amplitude
drop with rotational frequency is around 6% for the true pressure
amplitude. The decrement of amplitude depends on the tip speed
ratio , and the error is 8% at = 6.5, but a tendency of a pressure
change can be measured. In addition, the phase delay gets larger
as the frequency increases. The phase delay becomes around
25 deg at the optimal tip speed ratio. The scanner conversion time
is 3.2 ms, so during a pressure scan the azimuth changes by

Table 1 Geometrical specification of the test blade

Axial Force Ca
Lift Cl

r/R

Chord length
m

Twist angle
deg

Airfoil

0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0

0.0700
0.1474
0.1396
0.1318
0.1240
0.1162
0.1084
0.1006
0.0928
0.0850

18.33
12.00
8.33
5.00
4.68
2.86
1.44
0.91
0.00

Circle
DU91-W2-250
DU91-W2-250
DU93-W-210
DU93-W-210
NACA63-618
NACA63-618
NACA63-618
NACA63-215
NACA63-215

Drag Cd

Chord Line

Rotational Force
Cr

 + 


Rotating Plane

Fig. 4 Aerodynamic forces coefficients

031018-2 / Vol. 130, AUGUST 2008

Transactions of the ASME

Downloaded 06 Dec 2011 to 193.204.249.101. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

 =0

0.5

 =+15

0.8

Power Coefficient Cpower

Normalized amplitude

1.0

0.6
0.4
0.2
0
120

 =-15

0.4

 =+30
0.3

 =-30
 =+45

0.2

 =-45

0.1

Phase-lag [deg]

100

80
60

(a)

40

 =0

0.10

 =+15
0

10

20
30
Frequency [Hz]

40

50

Fig. 5 Dynamic response of pressure tubing

5.6 deg. In order to measure at the correct azimuth, the trigger


signal of the pressure sensor was set at 30 deg. Additionally, if the
tip speed ratio changes, an error of occurs. This error is 19 deg
at = 2.0 and +4 deg at = 6.2.

Results and Discussion

5.1 Power Curve. Figure 6a shows the variation of the output power coefficient Cpower with respect to the tip speed ratio for
various yaw angles. As the yaw angle increases, the inflow energy
through the rotor disk will decrease. However, when the actual
wind turbine is operating under field conditions, yaw is caused by
a change in wind direction, and it is not a thing to set artificially.
Therefore, in this study, Cpower is defined based on the rotor swept
area and the mainstream wind velocity. In addition, the tip speed
ratio is defined as the ratio of the blade tip speed versus the
mainstream wind velocity.
The output performance for = 0 has the expected form. In the
region 5.2, Cpower increases as decreases. The power coefficient reaches its peak value Cpower = 0.450 at = 5.2. By decreasing , the angle of the inflow will increase. Therefore, at a certain
point, stall will appear, which explains the sudden drop of Cpower.
Of course, the power output is subjected to induction, which finally determines the effective angle of attack as well to the relative velocity. As decreases, these two effects are competing so
that there will be an optimum. In yawed conditions, as increases, maximum Cpower decreases, and so does the optimum .
At approximately = 3.8, as decreases, Cpower for = 0 suddenly drops, but at higher yaw angles, the decrease of Cpower is
reduced. The fact that the optimum moves at lower as the yaw
angle increases is because the normal to the rotor disk velocity is
proportional to cos . Furthermore, in the low tip speed ratio
region, Cpower rises with increasing . In addition, since there are
no ground effects and the inflow is uniform, the flow will be
symmetric and therefore there will be no difference in Cpower between a plus yaw angle and minus one.
Figure 7 shows the relative maximum power as function of the
yaw angle defined as the ratio of the maximum power coefficient
Cpower for each yaw angle to the maximum for = 0. A curve of
cos2 6 and cos3 7 are shown in the same figure. At optiJournal of Solar Energy Engineering

 =-15

0.08
0.06
0.04






=+30
=-30
=+45
=-45

4
Tip Speed Ratio 

0.02
0

 -Ctorque curve

(b)

Fig. 6 Rotor performance distributions for various yaw


angles: a -Cpower curve; b -Ctorque curve

mum operation, it is expected that the relative velocity to the


blade element should have the good direction to get the optimum
angle of attack. Therefore, one would expect that the change is
proportional to the change of the axial inflow velocity. This means
that for the power, one would expect a cos3 dependence at first
order. When the absolute of increases, maximum Cpower decreases. The measured maximum Cpower for each are slightly
higher than cos3 with increasing . It seems that the difference
between this measurement and cos3 is caused by the effect of
complex variation on the sectional aerodynamic force with the
azimuth angle.

1.0
0.9
C power /Cpower 0

Torque Coefficient Ctorque

20

4
Tip Speed Ratio 
 -Cpower curve

0.8
0.7
0.6

Experiment

0.5

co s 

0.4

co s 

0.3
-45

2
3

-30

-15
0
+15
Yaw Angle  [deg]

+30

+45

Fig. 7 Comparison of maximum power coefficients for various


yaw angles

AUGUST 2008, Vol. 130 / 031018-3

Downloaded 06 Dec 2011 to 193.204.249.101. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Rotational Force Coefficient Cr

0.4

0.2

0.0
=0
=+15

-0.2
Fig. 8 Comparison of the pressure distribution at = 0 for
various yaw angles r / R = 0.7

4
Tip Speed Ratio 

=+30
 =+45
6

 -Cr curve (=0)

(a)

= tan1cos /r/R + sin cos

 =+15
 =+30
 =+45

1.0

(b)

4
Tip Speed Ratio 
 -Ca curve (=0)

Fig. 10 Rotational and axial force coefficients at = 0 for various yaw angles r / R = 0.7: a -Cr curve = 0; b -Ca curve
= 0

is small.
Figures 10a and 10b show the -Cr and -Ca curves for
= 0. Both figures correspond to r / R = 0.7. When = 0, and
4.0 as decreases, Cr increases and then reaches a maximum
value of Cr = 0.30 at = 4.0. In the region of 4.0, Cr drops
abruptly at = 3.8. The = 3.8 almost agrees with the at which
rotor torque drops in Fig. 6b. From Fig. 10b, when = 0, Ca
increases gently as decreases in the region of 4.5, and Ca
becomes approximately constant in the vicinity of = 4 4.5. Ca
suddenly drops in the vicinity of = 3.8, in agreement with the
value of corresponding to a sudden drop in Cr. It appears that
the reason is that the angle of attack to the blade element is large,
and the flow on the suction side is separated Fig. 11.

-6

25

 =0

20
15

 =+15
 =+30

10
5
0

 =0

1.5

0.5

Pressure Coefficien Cp

GeometricalAngleofAttack[deg]

Figure 9 shows the relationship between the geometric angle of


attack Eq. 1 and azimuth angle at r / R = 0.7, = 3.8, for example. In the case of yawed operation, the geometrical angle of
attack shows a minimum at = 0 and a maximum at = 180. The
angle of attack at = 0 is reduced as increases. The angle of
attack at = 180 is dependent on and the quantity of the change

Axial Force Coefficient Ca

2.0

5.2 Curves of Rotational Force Coefficient cr and Axial


Force Coefficient ca. In this section, the dependence of the axial
and rotational force coefficients Ca and Cr on the tip speed ratio
and the yaw angle are investigated at one radial position r / R
= 0.7 and four azimuth positions of 0 deg, 90 deg, 180 deg, and
270 deg.
Figure 8 shows the pressure distributions at r / R = 0.7 and
= 0. The pressure coefficient C p is shown against the chordwise
position x / c. C p is calculated from the surface pressure divided by
the local dynamic pressure. Local velocity is defined as the geometrical relative velocity based on the wind speed, the local blade
speed, , and . The C p defined with respect to the local inflow
velocity filters the change of the local speed with azimuth angle.
The tip speed ratio at which Ctorque reaches its maximum is
= 4.7, = 0. The pressure distribution shows leading edge suction
peak, so the blade element produces a large lift force. The suction
peak at the leading edge decreases as increases. The reason for
this is that the angle of attack for the blade element decreases as a
result of an increasing . It is caused by the increase of the
tangential component and the decrease of the axial component of
the blade element at = 0.
While the effective angle of attack is not determined by this
experiment directly, a geometrical angle of attack can be calculated from the velocity triangle defined by the wind velocity and
the rotational speed.

 =+45

90

180
270
Azimuth Angle  [deg]

360

Fig. 9 Geometrical angle of attack of r / R = 0.7 for various yaw


angles = 3.8

031018-4 / Vol. 130, AUGUST 2008

 = 1.8
 = 3.0
 = 4.0

-4

-2

2
0

0.2

0.4
0.6
Chord Station x/c

0.8

Fig. 11 Pressure distribution for various at = 0 r / R = 0.7

Transactions of the ASME

Downloaded 06 Dec 2011 to 193.204.249.101. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

0.4
Rotational Force Coefficient Cr

Rotational Force Coefficient Cr

0.4

0.2

 =0

0.0

- 0.2

 =+15
 =+30
 =+45

(a)

4
Tip Speed Ratio 

0.2

-0.2

 =0

Axial Force Coefficient Ca

Axial Force Coefficient Ca

4
Tip Speed Ratio 
 -Cr curve (=90)

2.0

2.0

(b)

 =+1 5
 =+3 0
 =+4 5

(a)

 -Cr curve (=180)

 =+15

1.5

 =+30
 =+45

1.0

0.5

 =0

0.0

4
Tip Speed Ratio 

 -Ca curve (=180)

(b)

 =0
 =+15
 =+30

1.5

 =+45

1.0

0.5

4
Tip Speed Ratio 

 -Ca curve (=90)

Fig. 12 Rotational and axial force coefficients at = 180 for


various yaw angles r / R = 0.7: a -Cr curve = 180; b -Ca
curve = 180

Fig. 13 Rotational and axial force coefficients at = 90 for various yaw angles r / R = 0.7: a -Cr curve = 90; b -Ca
curve = 90

From Figs. 10a and 10b, Cr and Ca at = 0 in the high tip


speed ratio region 4 decrease as increases in comparison
with = 0. When increases, the angle of attack decreases. The
value of showing a sudden drop in Cr and Ca results in a lower
tip speed ratio as increases. At r / R = 0.7, this is linked to the
reduction of the angle of attack for increasing . Additionally, the
maxima, Cr and Ca remain approximately constant and independent of the yaw angle, so it follows that the relationship between
the angle of attack of the blade element and changes according
to , but the relationship between the angle of attack and the Cr,
Ca produced by blade element does not change.
Figures 12a and 12b show the curves of -Cr and -Ca at
= 180, respectively. As increases, in the region of = 4, Cr
decreases slightly while in the region of 4, Cr increases. Also,
the maximum value of Cr is approximately the same for any . In
addition, for = 180, Cr exhibits an abrupt drop at = 4.0 except
for = 45. In Fig. 12b, Ca almost accords when = 5.0. Ca for
= 30 and 45 takes large values at low tip speeds due to the
blade-tower interaction. At such high yaw angles and at = 180,
the blade enters the region in which the tower is accelerating the
inflow. The effect of interference becomes large when the rotational speed is low. The value of at which Ca undergoes a
sudden drop for = 0 and 15 is almost the same. The reason is
that the in plane velocity due to yaw will be now opposite compared to the case shown in Fig. 10. Therefore, as increases, the
change on the angle of attack provoked by the decreasing axial
inflow velocity is counteracted by the increase of the in plane
wind velocity component resulting a much smaller shift.
Figures 13a and 13b show the -Cr and -Ca curves at
= 90, and Figs. 14a and 14b and show the same curves at
= 270. In Fig. 13a at = 90, the value of at maximum Cr

decreases while the maximum Cr increases with increasing . Ca


varies little with respect to when 4.0 in the case of = 90
Fig. 13b. At = 270, Cr decreases as increases in the region
of 4.0. The value corresponding to maximum Cr changes
with without a clear tendency. In addition, as increases, the
maximum Cr decreases. At = 270, Ca decreases as increases
when 4.0 Fig. 14b. The geometrical angle of attack at
= 90 becomes equal to that at = 270, but -Cr and -Ca are
largely different. A possible explanation is given below. First, only
the region of 4.0 is considered. At = 90, the in plane velocity
component of the inflow is in the direction from blade tip to root.
In general, the flow around the wind turbine expands from root to
tip because the flow speed is slowed down by the energy extraction of the rotor disk. A radial inflow velocity component pointing
toward the root prevents this. On the contrary, at = 270, the
radial inflow velocity will point toward the tip, i.e., in the same
direction as the radial direction velocity component of the expanded flow, and thus the resultant radial flow will increase. This
difference in the radial flow component on the blade at = 90 and
= 270 would lead to different aerodynamic loads.
Furthermore, relative flow velocity in the region of 4.0, Cr
at = 180 is small in comparison to its maximum, and the flow is
thought to separate from the airfoil Fig. 15. It is known that in
dynamic stall the -Cl curve exhibits hysteresis leading to retarded reattachment 13. So although the geometrical angle of
attack at = 270 decreases with respect to that at = 180, the flow
remains separated. In addition, the complicated variation of Cr at
= 270 with respect to is thought to be caused by the fact that
the hysteresis loop depends on the characteristics of the airfoil, the
nondimensional frequency, and the amplitude of the angle of attack variation.

Journal of Solar Energy Engineering

AUGUST 2008, Vol. 130 / 031018-5

Downloaded 06 Dec 2011 to 193.204.249.101. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Rotational Force Coefficient Cr

1 Increasing the yaw angle, the maximum power coefficient


decreases while the optimum tip speed ratio is lower.
2 The power coefficient of the rotor defined with respect to
the main stream wind velocity becomes higher when the
yaw angle is larger at low tip speed operation. The reason is
that there is an optimal value of angle of attack when the
blade is moving forward with respect to the main flow under yawed conditions so the blade lift coefficient is high.
3 Under yaw at r / R = 0.7, the blade passing from = 90
greatly contributes to the rotor torque. However, the blade
placed opposite has a smaller contribution.
4 Under yaw at r / R = 0.7, Ca and Cr are modulated by the
change in the geometrical angle of attack at = 0 and 180.
However, at = 90 and 270, the coefficients differ although
the geometrical angle of attack is the same. In attached flow
conditions high values, the aerodynamic performance is
thought to be affected by the radial velocity component and
the wake induced reduction in addition to the geometrical
change in the angle of attack. Pressure data do not allow to
confirm this argument as this would require velocity measurements.
5 In the low region, the blade at = 270 contributes less to
rotor torque as compared to that at = 90. It is thought that
the flow separation, which occurs around = 180 causing
the drop in Cr, has not reattached as a result of hysteresis.

0.4

0.2

0.0

-0.2
0

=+30
=+45

=0
=+15
2
4
Tip Speed Ratio

Axial Force Coefficient Ca

(a) -Cr curve (=270)

2.0

=0
=+15
=+30
=+45

1.5

Nomenclature
c
Ca
Cd
Cl
Cp

1.0

0.5
0

4
Tip Speed Ratio

(b) -Ca curve (=270)


Fig. 14 Rotational and axial force coefficients at = 270 for
various yaw angles r / R = 0.7: a -Cr curve = 270; b -Ca
curve = 270

Conclusions

Pressure data on a rotating blade at different yaw angles have


been obtained and analysis has been carried out with respect to the
resulting performance characteristics. The findings can be summarized as follows.

Cpower
Ctorque
Cr
r
R
T
U0
x

local chord length m


axial force coefficient
drag force coefficient
lift force coefficient
pressure coefficient
power coefficient= T / 1 / 2R2U30
torque coefficient= T / 1 / 2R3U20
driving force coefficient
radial position m
rotor radius= 1.2 m
rotor torque N m
uniform wind speed m/s
chordwise length from blade leading edge m
angle of attack deg
blade pitch deg
tip speed ratio= R / U0
blade twist angle deg
density of air kg/ m3
rotor speed 1/s
azimuth angle deg
yaw angle deg

References

Pressure Coefficient Cp

-6
 =0
 = 90
 = 180
 = 270

-4

-2
0
2

0.2

0.4
0.6
Chord Station x/c

0.8

Fig. 15 Pressure distribution for various at = 30 r / R


= 0.7

031018-6 / Vol. 130, AUGUST 2008

1 Snel, H., 1998, Review of the Present Status of Rotor Aerodynamics, Wind
Energy, 1S1, pp. 4669.
2 Vermeer, L. J., Srensen, J. N., and Crespo, A., 2003, Wind Turbine Wake
Aerodynamics, Prog. Aerosp. Sci., 3967, pp. 467510.
3 Schepers, J. G., Brand, A. J., Bruining, A., Graham, J. M. R., Hand, M. M.,
Infield, D. G., Madsen, H. A., Maeda, T., Paynter, J. H., van Rooij, R.,
Shimizu, Y., Simms, D. A., and Stefanatos, N., 2002, Final Report of IEA
Annex XVIII: Enhanced Field Rotor Aerodynamics Database, ECN-C-02016, p. 353.
4 Maeda, T., Ismaili, E., Kawabuchi, H., and Kamada, Y., 2005, Surface Pressure Distribution on a Blade of a 10 m Diameter HAWT Field Measurements
Versus Wind Tunnel Measurements, ASME J. Sol. Energy Eng., 1272, pp.
185191.
5 Simms, D., Schreck, S., Hand, M., and Fingersh, L. J., 2001, NREL Unsteady
Aerodynamics Experiment in the NASA-Ames Wind Tunnel: A Comparison of
Predictions to Measurements, NREL/TP-500-29494.
6 Imamura, H., Takezaki, D., Hasegawa, Y., Kikuyama, K., and Kobayashi, K.,
2004, Numerical Analysis of a Local Angle of Attack to HAWT Rotor Blade
in Unsteady Flow Conditions, Proceedings of European Wind Energy Conference & Exhibition 2004, London, UK, CD-ROM, p. 8.
7 Pesmajoglou, S., and Graham, J. M. R., 1993, Prediction of Yaw Loads on a

Transactions of the ASME

Downloaded 06 Dec 2011 to 193.204.249.101. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Horizontal Axis Wind Turbine, Proceedings of European Community Wind


Energy Conference, Lbeck-Travemnde, Germany, pp. 420423.
8 Haans, W., Sant, T., van Kuik, G., and van Bussel, G., 2005, Measurement of
Tip Vortex Paths in the Wake of a HAWT Under Yawed Flow Conditions,
ASME J. Sol. Energy Eng., 1274, pp. 456463.
9 Medici, D., and Alfredsson, P. H., 2006, Measurements on a Wind Turbine
Wake: 3D Effects and Bluff Body Vortex Shedding, Wind Energy, 93, pp.
219236.
10 Grant, I., Parkin, P., and Wang, X., 1997, Optical Vortex Tracking Studies of
a Horizontal Axis Wind Turbine in Yaw Using Laser-Sheet, Flow Visualization, Exp. Fluids, 23, pp. 513519.

Journal of Solar Energy Engineering

11 Maeda, T., Kamada, Y., Sakai, Y., and Takahara, N., 2005, Experimental
Study on Flow Around Blades of Horizontal Axis Wind Turbine in Wind
Tunnel, Trans. Jpn. Soc. Mech. Eng., Ser. B, 71701, pp. 171176.
12 Maeda, T., Kamada, Y., Sakai, Y., and Takahara, N., 2005, Experimental
Study on Flow Around Blades of Horizontal Axis Wind Turbine in Wind
Tunnel Second Report Studies on the Flow Around Blade Based on Pressure
Distribution, Trans. Jpn. Soc. Mech. Eng., Ser. B, 71705, pp. 13831389.
13 Amandolse, X., and Szchnyi, E., 2004, Experimental Study of the Effect
of Turbulence on a Section Model Blade Oscillating in Stall, Wind Energy,
74, pp. 267282.

AUGUST 2008, Vol. 130 / 031018-7

Downloaded 06 Dec 2011 to 193.204.249.101. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

You might also like