ArticlePDF Available

Reduction of Wind Turbine Noise Using Optimized Airfoils and Trailing-Edge Serrations

Authors:
  • United Technologies Corporation - Aerospace Systems

Abstract and Figures

Acoustic field measurements were carried out on a 94-m-diam three-bladed wind turbine with one standard blade, one blade with trailing-edge serrations, and one blade with an optimized airfoil shape. A large horizontal microphone array, positioned at a distance of about one rotor diameter from the turbine, was used to locate and quantify the noise sources in the rotor plane and on the individual blades. The acoustic source maps show that for an observer at the array position, the dominant source for the baseline blade is trailing-edge noise from the blade outboard region. Because of convective amplification and directivity, practically all of this noise is produced during the downward movement of the blade, which causes the typical swishing noise during the passage of the blades. Both modified blades show a significant trailing-edge noise reduction at low frequencies, which is more prominent for the serrated blade. However, the modified blades also show tip noise at high frequencies, which is mainly radiated during the upward part of the revolution and is most important at low wind speeds due to high tip loading. Nevertheless, average overall noise reductions of 0.5 and 3.2 dB are obtained for the optimized blade and the serrated blade, respectively.
Content may be subject to copyright.
UNCLASSIFIED
Executive summary
UNCLASSIFIED
Nationaal Lucht- en Ruimtevaartlaboratorium
National Aerospace Laboratory NLR
This report is based on an article published in AIAA Journal, Vol. 47, No. 6, June 2009.
Report no.
NLR-TP-2009-401
Author(s)
S. Oerlemans
M. Fisher
T. Maeder
K. Kögler
Report classification
UNCLASSIFIED
Date
August 2009
Knowledge area(s)
Aëro-akoestisch en experimenteel
aërodynamisch onderzoek
Descriptor(s)
wind turbines
microphone arrays
aerodynamic noise
Reduction of wind turbine noise using optimized airfoils and
trailing-edge serrations
Baseline SIROCCO Serration
Problem area
Wind turbine noise is a major
hindrance for the widespread use of
wind energy. For a modern large
wind turbine, aerodynamic noise
from the blades is the dominant
noise source.
Description of work
Acoustic field measurements were
carried out on a 94-m diameter
three-bladed wind turbine with one
standard blade, one blade with
trailing edge serrations, and one
blade with an optimized airfoil
shape. A large horizontal
microphone array was used to
locate and quantify the noise
sources in the rotor plane and on the
individual blades. The acoustic
performance of the different blades
was investigated as a function of
wind speed, observer position, and
blade azimuth.
Results and conclusions
Both modified blades show a
significant trailing-edge noise
reduction at low frequencies, which
is more prominent for the serrated
blade. However, the modified
blades also show tip noise at high
frequencies, which is mainly
radiated during the upward part of
the revolution and is most important
at low wind speeds due to high tip
loading. Nevertheless, average
overall noise reductions of 0.5 dB
and 3.2 dB are obtained for the
optimized blade and the serrated
blade, respectively.
Applicability
This study has demonstrated that
wind turbine noise can be halved
without adverse effects on the
aerodynamic performance.
UNCLASSIFIED
UNCLASSIFIED
Reduction of wind turbine noise using optimized airfoils and trailing-edge
serrations
Nationaal Lucht- en Ruimtevaartlaboratorium, National Aerospace Laboratory NLR
Anthony Fokkerweg 2, 1059 CM Amsterdam,
P.O. Box 90502, 1006 BM Amsterdam, The Netherlands
Tele
p
hone +31 20 511 31 13, Fax +31 20 511 32 10, Web site: www.nlr.nl
Nationaal Lucht- en Ruimtevaartlaboratorium
National Aerospace Laboratory NLR
NLR-TP-2009-401
Reduction of wind turbine noise using optimized
airfoils and trailing-edge serrations
S. Oerlemans, M. Fisher1, T. Maeder2 and K. Kögler1
1 GE Energy
2 GE Global Research
This report is based on an article published in AIAA Journal, Vol. 47, No. 6, June 2009.
The contents of this report may be cited on condition that full credit is given to NLR and the authors.
This publication has been refereed by the Advisory Committee AEROSPACE VEHICLES.
Customer NLR
Contract number ----
Owner NLR
Division NLR Aerospace Vehicles
Distribution Unlimited
Classification of title Unclassified
October 2009
Approved by:
Author
Reviewer Managing department
NLR-TP-2009-401
2
Contents
I Introduction 3
II Experimental Method 5
A Test Setup 5
B Data Acquisition 5
C Test program 6
D Phased-Array Processing 7
E Accuracy of Sources Localization and Quantification 7
III Results and Discussion 8
A Noise Sources in the Rotor Plane 8
B Noise Sources on the Individual Blades 9
IV Conclusions 13
References 13
Reduction of Wind Turbine Noise Using Optimized
Airfoils and Trailing-Edge Serrations
Stefan Oerlemans
National Aerospace Laboratory/NLR, 8300 AD Emmeloord, The Netherlands
Murray Fisher
GE Energy, Greenville, South Carolina 29615
Thierry Maeder
GE Global Research, 85748 Munich, Germany
and
Klaus Kögler§
GE Energy, 48499 Salzbergen, Germany
DOI: 10.2514/1.38888
Acoustic eld measurements were carried out on a 94-m-diam three-bladed wind turbine with one standard blade,
one blade with trailing-edge serrations, and one blade with an optimized airfoil shape. A large horizontal microphone
array, positioned at a distance of about one rotor diameter from the turbine, was used to locate and quantify the noise
sources in the rotor plane and on the individual blades. The acoustic source maps show that for an observer at the
array position, the dominant source for the baseline blade is trailing-edge noise from the blade outboard region.
Because of convective amplication and directivity, practically all of this noise is produced during the downward
movement of the blade, which causes the typical swishing noise during the passage of the blades. Both modied blades
show a signicant trailing-edge noise reduction at low frequencies, which is more prominent for the serrated blade.
However, the modied blades also show tip noise at high frequencies, which is mainly radiated during the upward
part of the revolution and is most important at low wind speeds due to high tip loading. Nevertheless, average overall
noise reductions of 0.5 and 3.2 dB are obtained for the optimized blade and the serrated blade, respectively.
Nomenclature
D= trailing-edge noise directivity function
f= frequency
M= local blade inow Mach number
N= number of measurements
P= rotor power
St = Strouhal number (f=U)
U= local blade inow velocity
U10 = wind speed at 10 m height
= misalignment angle between array and wind turbine
= trailing-edge boundary-layer displacement thickness
"= standard deviation of the mean (standard error) (= 
N
p)
= angle between blade chord line and source-observer line
= angle between blade ow velocity and source-observer
line
= standard deviation
= angle between blade plane and plane containing chord
line and observer
= blade azimuth angle
I. Introduction
WIND turbine noise is one of the major issues for the wide-
spread use of wind energy. For a modern large wind turbine,
aerodynamic noise from the blades is generally considered to be the
dominant noise source, provided that mechanical noise is adequately
treated [1]. The sources of aerodynamic noise can be divided into
airfoil self-noise and inow-turbulence noise. Airfoil self-noise is the
noise produced by the blade in an undisturbed inow and is caused
by the interaction between the boundary layer and the trailing edge of
the blade. Self-noise can be tonal or broadband in character and may
be caused by several mechanisms, such as turbulent boundary-layer/
trailing-edge interaction noise (subsequently denoted as trailing-
edge noise), laminar boundary-layer vortex-shedding noise, trailing-
edge bluntness noise, or blade tip noise. Inow-turbulence noise is
caused by the interaction of upstream atmospheric turbulence with
the blade and depends on the atmospheric conditions. It is an open
issue to what extent inow-turbulence noise contributes to the
overall sound level of a wind turbine [2].
Because of the large number of applications (e.g., wind turbines,
airplanes, helicopters, and fans), the characteristics of airfoil noise
have been investigated extensively in both experimental and theo-
retical studies [313]. Both inow-turbulence and self-noise mech-
anisms were considered and the dependence on parameters such as
ow speed, angle of attack, radiation direction, and airfoil shape
was characterized. These studies formed the basis of several
semi-empirical wind turbine noise prediction models, which were
validated by comparison with eld measurements [1420]. Because
the eld results only provided the overall sound level of the turbine,
the relative importance of the different mechanisms was determined
mainly on the basis of the predictions. In some studies, inow-
turbulence noise was regarded to be the dominant source [11,14
16,18], whereas others considered trailing-edge noise to be dominant
[17]. In another case, the turbine noise in different frequency ranges
was attributed to mechanical noise, trailing-edge noise, tip noise, and
inow-turbulence noise [19].
Presented as Paper 2819 at the 14th AIAA/CEAS Aeroacoustics
Conference, Vancouver, Canada, 57 May 2008; received 2 June 2008;
revision received 23 February 2009; accepted for publication 24 February
2009. Copyright © 2009 by the National Aerospace Laboratory/NLR.
Published by the American Institute of Aeronautics and Astronautics, Inc.,
with permission. Copies of this paper may be made for personal or internal
use, on condition that the copier pay the $10.00 per-copy fee to the Copyright
Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923; include
the code 0001-1452/09 $10.00 in correspondence with the CCC.
Research Engineer, Aeroacoustics Group, P.O. Box 153.
Research Engineer, Advanced Aero & Acoustics Technologies, 300
Garlington Road.
Research Engineer, Alternative Energy and Environmental Technologies,
Freisinger Landstrasse 50.
§Research Engineer, Advanced Aero & Acoustics Technologies,
Holsterfeld 16.
AIAA JOURNAL
Vol. 47, No. 6, June 2009
1470
NLR-TP-2009-401
3
In a few studies, source location measurements were performed to
provide more direct information on the source mechanisms [2125].
The results from [2123] were obtained using an acoustic parabola or
a linear array of microphones and focused only on the horizontal
(downward) blade position ( 90 deg). In [24,25], a large two-
dimensional microphone array, positioned on the ground about one
rotor diameter upwind of the turbine, was used to localize the noise
sources in the complete rotor plane and on the individual blades for
two different wind turbines. It was shown that practically all recorded
noise was produced during the downward movement of the blades.
This strongly asymmetric source pattern, which causes the typical
swishing noise during the passage of the blades, was explained by
convective amplication and trailing-edge noise directivity. The
following directivity function for trailing-edge noise was used [26]:
D2sin2=2sin2
1Mcos 4(1)
where is the angle between the blade chord line and the source-
observer line, is the angle between the plane of the blade and the
plane containing the chord line and the observer, is the angle
between the (inverted) local blade inow velocity and the source-
observer line, and Mis the local blade inow Mach number (see
Fig. 1 for denition of angles). The numerator in Eq. (1) describes the
directivity of high-frequency trailing-edge noise and indicates that
most of the noise is radiated in the direction of the airfoil leading
edge. It was analytically derived for edge noise from a semi-innite
at plate [6,27], but was also found to be valid for nite airfoils [9],
provided that the angle is not too close to 180 deg. In the limit for
low-frequency dipole noise, for which the acoustic wavelength is
much larger than the airfoil chord, the sin2=2term changes into
sin2[5,26]. Nevertheless, in [24], Eq. (1) was found to be valid for
the whole tested frequency range, including the low frequencies, for
which the acoustic wavelength was of the same order as the blade
chord. The denominator in Eq. (1) represents the fourth-power
convective amplication factor for trailing-edge noise [28] and
indicates that the noise source becomes louder when it is moving
toward the observer.
Many studies have addressed the reduction of airfoil or wind
turbine noise. Because inow-turbulence noise and trailing-edge
noise both scale with approximately the fth power of the local blade
inow velocity [5,6,24], an obvious means for noise reduction is to
reduce the rotor rpm or rotor diameter. However, these measures also
reduce the power output of the turbine [1]. The same holds for
increasing the blade pitch angle (i.e., turning the blade leading edge
against the wind): this reduces the local angle of attack and therefore
the noise, but (due to the reduced lift) also the power production.
Thus, the challenge is to achieve a noise reduction without a
reduction in power output. With regard to tip noise, which depends
on the characteristics of the tip vortex, it has been demonstrated in
several studies that the tip shape can have a signicant inuence on
the noise from a wind turbine [1,6]. The importance of inow-
turbulence noise depends on the structure of the atmospheric
turbulence and on the shape of the blades. It has been shown both
experimentally and numerically that inow-turbulence noise levels
increase with increasing sharpness of the airfoil leading edge [10,13].
With regard to trailing-edge noise, a number of reduction concepts
have been investigated. After it had been shown theoretically that the
acoustic radiation efciency of a trailing edge can be reduced by
serrations [29] (see Fig. 2), this concept was investigated in a number
of experimental studies on 2-D airfoils [30], model wind turbines
[31,32], and a full-scale wind turbine [23]. In [31], serrations were
applied to a 16-m-diam model wind turbine and, depending on the
ow conditions, overall noise reductions of up to 3.5 dB were
obtained. To prevent increased noise at high frequencies, it was
found to be critical to align the plane of the serrations with the
trailing-edge ow. In [23], serrations were applied to a 1 MW wind
turbine, and an overall noise reduction of 23 dB was obtained,
despite increased noise at high frequencies. It should be noted that the
measurements in [23,31] only focused on the horizontal (downward)
blade position ( 90 deg), which may give an incomplete picture.
An alternative concept for trailing-edge noise reduction is the
application of exible trailing-edge brushes. The brushes align
automatically with the trailing-edge ow and have shown signicant
noise-reduction potential in wind-tunnel tests on at plates and on a
2-D airfoil [12,33]. However, a rst attempt to apply this concept to a
full-scale wind turbine yielded a reduction of only 0.5 dB [34],
possibly because the improvised brushes were too short. Finally, it
has been shown in calculations and wind-tunnel tests on 2-D airfoils
that trailing-edge noise can be reduced by a modication of the airfoil
shape, without a loss in aerodynamic performance [35]. Note that in
the case of an acoustically optimized airfoil shape, trailing-edge
noise is reduced by changing the structure of the boundary-layer
turbulence, whereas serrations or brushes are meant to affect only the
scattering at the trailing edge. Thus, the effects of an optimized airfoil
shape and brushes or serrations are expected to supplement each
other.
source
observe
r
ϕ
θ
-U
ξ
trailing edge
chord line
Fig. 1 Denition of angles for trailing-edge noise source.
Fig. 2 Climber removing trips from serrated blade.
OERLEMANS ET AL. 1471
NLR-TP-2009-401
4
The present study concerns acoustic eld measurements on a
2.3 MW, 94-m-diam wind turbine with one standard (baseline)
blade, one blade with an acoustically optimized airfoil shape, and one
standard blade with trailing-edge serrations. The tests were
performed in the framework of the European SIROCCO (Silent
Rotors by Acoustic Optimisation) project [34]. Building on the
results from earlier wind-tunnel studies on a model rotor [32], the
subject of the project was the design, testing, and full-scale validation
of quiet wind turbine blades, without a loss in power performance. In
an earlier stage of the project, acoustic eld measurements on the
baseline turbine [25] indicated that trailing-edge noise from the outer
25% of the blades was the dominant noise source for this turbine and
that the three blades produced practically the same sound pressure
levels: the average overall sound pressure level (OASPL) for the
three blades differed by less than 0.15 dB and, for the two standard
blades that were used again in the present campaign, less than
0.05 dB.
Subsequently, optimized airfoil shapes were designed and as-
sessed through aerodynamic and acoustic wind-tunnel tests on 2-D
airfoils [3436]. The principle of the airfoil design was to reduce the
dominant low-frequency (less than 1 kHz) trailing-edge noise peak
in the spectrum (which is due to the thick suction-side boundary
layer) by reducing the loading of the suction side, at the expense of
an increased pressure-side loading (which causes a slightly higher
noise level at less important medium frequencies of 13 kHz) [35].
The wind-tunnel tests showed 23 dB reduction in OASPL
(depending on lift coefcient) [34] and an improved aerodynamic
performance for the newly designed airfoils, even though severe
geometric and aerodynamic constraints had to be considered in the
design (to enable implementation in the existing blade structure).
The new airfoil was then incorporated in the design of the outer part
of the optimized blade (subsequently denoted as the SIROCCO
blade). The twist distribution of the SIROCCO blade was modied
such that the lift distribution was approximately the same as for the
baseline blade. In addition to the new blade design, the third rotor
blade was used to test a second noise-reduction concept, trailing-
edge serrations. The serrated blade had the same nominal geometry
as the baseline blade (including the twist distribution). From power
and loads measurements on the baseline and modied rotors, it was
found that the in-plane and out-of-plane blade loads on the serrated
blade (and, to a lesser extent, also on the SIROCCO blade) were
slightly higher than on the baseline blade, causing the aerodynamic
performance of the modied blades to be similar or slightly better
than the baseline blade [34].
The main goal of the present test campaign was to assess the
acoustic performance of the two modied blades versus the baseline
blade. To assess the effect of blade roughness due to dirt or insects,
the blades were tested in both clean and tripped conditions. A large
horizontal microphone array, positioned at a distance of about one
rotor diameter from the turbine, was used to measure the distribution
of the noise sources in the rotor plane and on the individual blades.
Because the array position was xed and the wind direction varied,
both up- and downwind measurements were performed. In the
present paper, the acoustic array results are presented and analyzed.
The noise characteristics for the three blades are investigated as a
function of wind speed, rotor azimuth angle, and observer position
(upwind or downwind), for clean and tripped conditions. Section II
describes the test setup, test program, and the array processing
methods. In Sec. III, the results are presented and discussed. The
conclusions of this study are summarized in Sec. IV.
II. Experimental Method
A. Test Setup
The measurements were carried out in March/April 2007 on the
same General Electric 2.3 MW prototype test wind turbine that was
used in the baseline test campaign in 2005. It had a rotor diameter of
94 m, a tower height of 100 m, and was located on the Netherlands
Energy Research Foundation test site in the Wieringermeer (The
Netherlands). The turbine control system adjusted the rpm and blade
pitch angle, depending on the wind speed measured at the nacelle: for
higher wind speeds, the pitch angle was increased, reducing the local
angle of attack and thus the blade loading. The rpm increased up to a
certain wind speed, after which it remained constant. The turbine had
a yaw mechanism that automatically turned the rotor against the
wind.
To compare the blade performance for identical weather and
turbine conditions, the rotor consisted of one standard (baseline)
blade, one standard blade with trailing-edge serrations, and one
SIROCCO blade. The SIROCCO blade was nominally identical to
the baseline blade, except for the outer 30%, where it had a new
airfoil. The serrated blade had the same nominal geometry as the
baseline blade. The aluminum serrations, with a thickness of 2 mm,
were mounted to the outer 12.5 m of the blade on the pressure side.
The 2 mm step was smoothed using ller material over a few
centimeters of chord. Similar to [32], the length of the serrations was
about 20% of the local chord and varied as a function of radius: the
tooth length was about 10 cm at the tip and about 30 cm at the most
inboard position. A picture of the serrated blade is shown in Fig. 2.
Using templates for different radial stations, the plane of the
serrations was aligned with the ow direction at the blade trailing
edge (as determined from ow calculations). This trailing-edge ow
direction is constant in the variable-rpm region of the turbine. By
aligning the serrations with the ow, it was attempted to minimize
their aerodynamic impact and prevent increased high-frequency
noise by ow through the teeth. Before the acoustic measurements,
all blades were cleaned. To assess the effect of blade roughness due to
dirt or insects, the blades were tested with and without trips: in state 1
all blades were tripped and in state 2 all blades were clean. The 2-D
trips, with a thickness of about 0.4 mm and a width of 4 mm, were
installed close to the leading edge on both sides of the blade, from the
very tip to about half the blade span. The (variable) blade pitch angle
was the same for the three blades.
An acoustic array was used to locate and quantify the noise sources
on the rotor and on the individual blades. The acoustic array
consisted of 148 Panasonic WM-61 microphones, mounted on a
horizontal wooden platform of 16 18 m2, which was positioned at
a distance of about one rotor diameter from the turbine (Fig. 3).
Because the array position was xed and the wind direction varied,
both up- and downwind measurements were performed. The
Panasonic microphones were mounted ush to the surface of the
platform, with the membrane parallel to the platform, and were
equipped with wind screens. As a reference, two calibrated LinearX
M51 microphones equipped with hemisphere wind screens were
mounted on the platform as well. To correct for the view angle of
about 45 deg (Fig. 3), the microphone array had an elliptic shape
(Fig. 4), rather than the conventional round array design. In this way,
the effective array shape, as seen from the rotor, is round, so that the
resolution with which the noise sources in the rotor plane are
localized is approximately the same in the horizontal and vertical
directions. The ellipse was slightly tilted to the right-hand side of the
rotor plane, to obtain maximum resolution on the side where the
blades move downward [for the standard array position (i.e., upwind
of the turbine)] and where maximum noise radiation was observed in
the 2005 campaign. The array had a high microphone density in the
center to ensure low side-lobe levels at high frequencies and had a
low-density outer part to obtain a good resolution at low frequencies
[37]. The distance and orientation of the array with respect to the
turbine were determined using a laser distance meter and a compass.
B. Data Acquisition
Acoustic data from the array microphones were synchronously
measured using the VIPER multichannel data-acquisition system
[38] at a sample frequency of 30.7 kHz and a measurement time of
30 s. The acoustic data were processed using fast Fourier transform
blocks of 1024 samples with a Hanning window and 50% overlap,
yielding 1800 averages and a narrowband frequency resolution of
30 Hz. A second-order 500 Hz high-pass lter (12 dB=octave [38])
was used to suppress high-amplitude pressure uctuations at low
frequencies and thus to extend the dynamic range of the A/D
converter, so that low-pressure amplitudes at high frequencies were
1472 OERLEMANS ET AL.
NLR-TP-2009-401
5
included. The sound levels were corrected for the lter response and
for pressure doubling due to reections at the platform. Before the
measurements, the sensitivity at 1 kHz was determined for all
array microphones using a calibrated pistonphone. The frequency
response of the Panasonic microphones was taken from previous
calibration measurements. The frequency response of the M51
microphones was taken from calibration sheets. No corrections were
applied for microphone directivity, because calibration measure-
ments showed that these effects amounted to less than 2 dB up to
20 kHz for angles smaller than 75 deg with respect to the microphone
axis. Phase matching of the microphones was checked using a
calibration source at known positions. A trigger signal from the
turbine (one pulse per revolution) was recorded synchronously with
the acoustic data to determine the location of the blades as a function
of time for the source localization on the individual blades
(Sec. II.D).
In parallel to the acoustic measurements, several parameters from
the turbine and two nearby meteorological masts were continuously
measured at a sample rate of 4 Hz or higher. These data included wind
speed, wind direction, temperature, power production, turbine
orientation (misalignment angle ), rpm, and blade pitch angle.
C. Test Program
During the 4-week test campaign, in total, more than 600
measurements were taken. Following the International Electro-
technical Commission norm for wind turbine noise measurements
[39], it was attempted to obtain measurements at wind speeds (at
10 m height) between 6 and 10 m=s. The wind speed at 10 m height
was calculated by multiplying the average wind speed measured at
the nacelle by 0.70 (according to the standard wind prole from
[39]). Because the array position was xed and the wind direction
varied, both up- and downwind measurements were performed.
Measurements with a large misalignment angle (see Fig. 3) were
excluded from the analysis, because for very oblique view angles, the
array resolution becomes poor. On the basis of the turbine
operational data, the most stable measurements (i.e., small variation
in wind speed, rpm, pitch angle, and turbine orientation) were
selected for further analysis.
An overview of the selected measurements is given in Table 1.
Because of unpredictable weather conditions, it was not possible to
obtain measurements in each wind-speed bin, and for state 1, only
downwind measurements were done. Because the clean rotor is
considered to be most representative for the rotor during normal
operation, and because the upwind measurements covered all
relevant wind speeds, the focus of this paper will be on state 2a.
The average turbine and weather conditions for the different rotor
states are listed in Table 2 (equal weights per wind-speed bin). As
mentioned in Sec. II.A, the blade pitch angle (not to be confused with
the twist distribution) is zero at low wind speeds and increases for
higher wind speeds. To give an impression of the variation of the
parameters within each state, this table also shows the standard
deviation for each value dened as

1
N1X
N
i1xi
x2
v
u
u
t
Wind
Platform 45°
Tu rb i n e
ψ
Platform
Turbine
α
Wind
Turbine
Platform
Fig. 3 Side view (left), front view (middle), and top view (right) of the test setup. The array microphones were mounted on the platform in an elliptic
shape for optimum resolution.
Table 1 Number of selected measurements per wind-speed bin for each rotor state
6m=s7m=s8m=s9m=s10m=s
State 1 (tripped rotor; array downwind) 8 8 8 0 0
State 2a (clean rotor; array upwind) 7 8 8 8 6
State 2b (clean rotor; array downwind) 8 8 8 0 0
Table 2 Average weather and turbine parameters for each rotor statea
U10,m=sP, MW rpm , deg Blade pitch, deg
State 1 6.9 1.6 (0.1) 14.5 (0.2) 204 (12) 0.0 (0.1)
State 2a 8.1 2.1 (0.1) 14.9 (0.0) 2(4.4) 5.1 (0.6)
State 2b 6.9 1.6 (0.1) 14.6 (0.1) 183 (2.4) 0.0 (0.0)
aThe standard deviation is indicated between parentheses.
-10
-6
-2
2
6
10
-8 -4 0 4 8
Y [m]
X [m]
Fig. 4 Layout of array microphones. The rectangle indicates the
platform dimensions.
OERLEMANS ET AL. 1473
NLR-TP-2009-401
6
with
x1
NX
N
i1
xi
Note that the power, rpm, and blade pitch angle are not randomly
distributed around the mean value, but depend on the wind speed
according to the turbine control system. Therefore, the standard
deviations for these parameters are based on linear curve ts through
the measured data as a function of wind speed. Because the turbine
had an automatic yaw system, the yaw angle (i.e., the difference
between the wind direction and the turbine orientation) was assumed
to be zero.
D. Phased-Array Processing
The microphone array data were processed using two different
methods. With the rst (stationary) method, noise sources in the
complete rotor plane were localized using conventional beamform-
ing [40]. Thus, noise from the rotor hub can be separated from blade
noise, and it can be seen where in the rotor plane the blade noise is
produced. The method shows the integrated effect of the three blades,
averaged over the complete measurement time of 30 s (i.e., several
revolutions).
The rst step of this processing involves the calculation of an
averaged cross-spectral matrix, which contains the cross powers of
all microphone pairs in the array. To improve the resolution and to
suppress background noise (e.g., wind-induced pressure uctuations
on the microphones), the main diagonal of the cross-power matrix
(i.e., the auto powers) was discarded. A spatial window was applied
to the microphone signals, which reduced the effective array size
with increasing frequency and which corrected for the variation in
microphone density over the surface of the array [37]. In this way, the
array resolution at low frequencies was improved, and coherence-
loss effects at high frequencies (due to propagation of the sound
through the atmospheric boundary layer) were suppressed.
Acoustic source maps of the rotor plane were produced by
electronically steering the array to a set of grid points and calculating
the noise radiated from each of them. The scan grid, with a diameter
of 140 m and a mesh width of 2 m, was placed in the rotor plane and
rotated in accordance with the orientation of the turbine (depending
on wind direction). The 4 deg tilt angle between the rotor axis and the
horizontal plane was also taken into account. The effect of sound
convection in the atmospheric boundary layer was taken into account
by assuming a constant wind speed between the scan location and
the microphones. This constant wind speed was calculated as the
average wind speed between the rotor hub and the array center, as
determined from the standard wind velocity prole in [39].
The narrowband acoustic source maps were summed to one-third-
octave bands, and the source levels were normalized to a constant
reference distance. The noise sources in the rotor plane were
quantied using a source-power integration method [37]. This
technique sums the source powers in (part of) the measured source
map and corrects the results with a scaling factor obtained by
performing a simulation for a monopole source at the center of the
integration region. The thus-obtained integrated sound pressure level
of the turbine, as measured at the array position, is similar to the
apparent sound power level dened in [39]. All spectra presented in
this paper are in one-third-octave bands. The accuracy of the
integration method is discussed in the next section.
The second processing method employed three rotating scan
planes to localize the (de-Dopplerized) noise sources on the three
individual blades [rotating source identier (ROSI)] [41]. This
enabled a comparison of the noise from the different blades. The start
position of the scan planes was determined using a trigger signal from
the turbine that was recorded synchronously with the acoustic data.
Acoustic source maps of the different blades were produced by
electronically steering the array to a set of rotating grid points and
calculating the noise radiated from each of them. The three scan grids
were placed in the rotor plane at azimuthal positions corresponding
to the three blades. The blade grids ran from 15 to 60 m in the radial
direction, had a chordwise extent of 30 m, and had a mesh width of
1m.
Similar to the rst processing method, the narrowband acoustic
source maps were summed to one-third-octave bands, and the source
levels were normalized to a constant reference distance. The ROSI
results show the noise sources on the individual blades, averaged
over a specied part of the revolution. To distinguish between the
noise production during the down- and upward movements of the
blades, separate ROSI scans were done for blade azimuth angles
from 0 to 180 deg and from 180 to 360 deg (with 0 deg as the upper
vertical blade position). To limit processing time, only the rst
rotor revolution after the start of each acoustic measurement was
processed. The noise from the blades was quantied using a power
integration method for moving sound sources [42], which is similar
to the aforementioned integration method for the stationary rotor
plane. The thus-obtained integrated sound levels represent the
contribution of the different blades to the overall sound pressure level
of the turbine, as measured at the array position.
E. Accuracy of Source Localization and Quantication
The relative position and orientation of the acoustic array and the
wind turbine were determined using a laser distance meter and a
compass. Nevertheless, there are a number of uncertainties in the
localization method, which may cause a deviation between the
measured and actual source positions. First, the blades are not located
exactly in the rotor plane: the 4 deg rotor tilt angle is accounted for,
but not the rotor cone angle and the bending of the blades outside the
plane. Second, sound refraction by wind shear and sound convection
by wind gusts are not accounted for; a constant wind speed is
assumed. Third, the rotor rpm is assumed to be constant within
one revolution. Therefore, the accuracy of the source localization
technique was assessed by attaching a whistle to one of the blades for
a short period of time, at a position unknown to the acoustic test team.
After determining to which blade the whistle was mounted, the ROSI
source maps were used to estimate the exact whistle position (Fig. 5).
The thus-obtained source radius was found to deviate only 0.5 m
from the actual radius, which is considered to be accurate enough for
these tests. Figure 5 also illustrates that the scan resolution is
sufciently high to determine accurate, integrated, blade noise levels.
From numerical simulations [43], it was found that as long as the
distance between adjacent scan points is smaller than the main lobe
width (i.e., the width at 3 dB below the peak level), the integrated
levels are accurate.
The acoustic source maps were quantied using the power
integration method described in the previous section. The accuracy
of this method in terms of absolute sound pressure levels was veried
by comparing the integrated rotor source maps with the measured
sound pressure levels at the array microphones. If all the noise
measured by the array microphones is due to the turbine rotor, these
spectra should coincide. Figure 6 shows the spectra measured by the
array microphones (array) and the reference microphone at the center
of the array (refmic) versus the integrated spectra for the rotor
(powint) and the three blades (ROSI). These spectra were averaged
over all measurements in state 2a, were corrected for pressure
Fig. 5 Acoustic source map for whistle measurement. The black dots
indicate the scan grid and the cross indicates the actual whistle position.
1474 OERLEMANS ET AL.
NLR-TP-2009-401
7
doubling by the array surface, and were normalized to the same
reference distance. It can be seen that the average spectrum of the
array microphones is practically equal to that on the reference
microphone. The small difference at low and high frequencies may
be due to wind noise on the array microphones at the edge of the
platform.
Figure 6 also shows that the integrated spectra for the rotor and the
blades have the same shape as the reference spectrum, but are about
3 dB lower over the whole frequency range. This discrepancy cannot
be explained by the fact that the array method is applied to incoherent
extended sources, because simulations for an incoherent line source
yielded accurate integrated levels [37]. However, in addition to
the aforementioned possible deviation between the rotor plane
and the actual blade position, the observed difference may be
partly explained by certain assumptions and simplications in the
integration method, such as the use of a single monopole source at the
center of the integration region to determine the scaling factor for the
source powers. For a simulated, realistic, wind-turbine-rotor noise
source distribution, the difference between the actual and integrated
overall rotor noise level was about 1 dB [25]. The difference may also
be partly attributed to coherence loss at the array microphones, due
to propagation of the sound through the turbulent atmospheric
boundary layer. A similar effect has also been observed in open-jet
wind-tunnel tests [37]. However, coherence-loss effects typically
increase with frequency, whereas a more or less constant offset is
found here. Furthermore, coherence-loss effects would be expected
to increase with increasing wind speed, whereas the difference here
between the integrated rotor noise level and the level of the reference
microphone remained constant (within about 0.5 dB) for increasing
wind speed. An alternative explanation for the lower integrated
levels could be that the reference and array microphones pick up
some background noise (e.g., from the wind over the platform),
which is not present in the integrated rotor noise spectra (no
background noise measurements with a stopped rotor were per-
formed in the present test campaign).
The difference between the two integrated spectra may be due to
the fact that the ROSI spectrum is de-Dopplerized and the rotor
spectrum is not. For the down-going blade, de-Dopplerization results
in reduced frequencies and reduced sound levels [due to convective
amplication (see Sec. I)], and conversely for the up-going blade.
Furthermore, the different integration regions (complete rotor versus
outer part of the blades) will result in different scaling factors
(depending on, for example, the side-lobe behavior [37]), which may
lead to differences in the integrated spectra.
For the evaluation of the noise-reduction concepts in the present
study, the accuracy of the relative sound levels (i.e., level differences
between the different blades) was most important. This accuracy was
assessed in the baseline test campaign by comparing the individual
blade noise spectra for two consecutive revolutions: for each blade,
the overall sound pressure level (averaged over all selected
measurements) reproduced within 0.06 dB for the two consecutive
revolutions, and the differences between the blades reproduced
within 0.03 dB. It should be noted that due to the small difference in
the out-of-plane blade loads (see Sec. I), the bending may be different
for the three blades. Because the scan planes for all three blades are
placed in the rotor plane, this may affect the measured noise
differences between the blades. However, because the difference in
bending between the blades can be estimated to be quite small (less
than 0.25 m) on the basis of the load measurements and the spatial
array resolution perpendicular to the rotor plane is limited, the
systematical error in the noise difference can be determined to be less
than about 0.05 dB (from array simulations). This means that the
average level differences between the blades can be considered to be
accurate within 0.1 dB for the given weather conditions, turbine
operation parameters, and misalignment angle.
III. Results and Discussion
In this section, the results of the acoustic array measurements are
presented and discussed. First, the noise source distribution in the
rotor plane is analyzed for the up- and downwind array positions.
Next, the noise sources on the individual blades are investigated, to
assess the performance of the SIROCCO blade and the serrations as a
function of array position, wind speed, and rotor azimuth angle.
Because the clean rotor is considered to be most representative for the
rotor during normal operation, and because the upwind measure-
ments covered all relevant wind speeds, the focus will be on state 2a.
A. Noise Sources in the Rotor Plane
Each 30 s measurement resulted in acoustic source maps, showing
the noise sources in the rotor plane as a function of frequency. To
show the general trends, these maps were averaged over all selected
measurements in the respective rotor state (Figs. 79). Thus, these
maps show the average effect of all three blades over all revolutions.
The black circle indicates the 94-m rotor diameter and the X indicates
the center of the rotor plane. The range of the color scale is always
12 dB and the maximum is adjusted for each frequency band and
each rotor state. The purpose of these source maps is to show the
qualitative source characteristics; a quantitative comparison between
the different rotor states will be made in Sec. III.B. The rotation
direction is clockwise; note that the source maps in Figs. 7 and 9 are
mirrored to allow easy comparison with the upwind measurements
(i.e., an observer at the array position would see the rotor turn in the
counterclockwise direction). Similar to the previous results on the
baseline turbine [25], the upwind measurements (Fig. 8) indicate that
for an observer at the array position, most of the noise is produced by
160 315 630 1250 2500 5000
Frequency (Hz)
SPL (dBA)
array
refmic
ROSI
powint
5 dB
Fig. 6 Verication of power integrated method.
Fig. 7 Average stationary source maps for state 1 (tripped rotor,
downwind array position). The range of the color scale is 12 dB and the
maximum is adjusted for each frequency band.
OERLEMANS ET AL. 1475
NLR-TP-2009-401
8
the outer 25% of the blades during their downward movement. This
effect, which causes the typical swishing noise during the passage of
the blades, can be explained by convective amplication and trailing-
edge noise directivity [24], as described in Eq. (1). For the higher
frequencies, minor noise sources can also be observed at the tip of the
up-going blades and at the location where the blades pass the tower.
The nature of this tower source is hard to assess on the basis of the
present data, but it could originate from 1) reection of blade noise on
the tower, 2) impingement of blade tip vortices on the tower, and/or
3) the upstream inuence of the tower on the oweld around the
blade.
For the downwind measurements (Figs. 7 and 9), nacelle noise
appears to be more pronounced than for the upwind array position. In
these plots, the nacelle source appears offcenter, because it is located
in front of the scan plane, which coincides with the rotor plane. Only
about 2.5 dB of the observed difference between the up- and
downwind nacelle noise levels can be explained by the smaller
distance to the array and the distance between the source and the scan
plane. The remaining difference may be explained by a combination
of two factors: First, mechanical noise generated inside the nacelle is
radiated mainly in the downwind direction, because the ventilation
opening is on the rear side of the nacelle. The relative nacelle noise
level in state 1 is lower than in state 2b, probably because of the
misalignment angle of 204 deg (i.e., a deviation of 24 deg with
respect to 180 deg) in state 1 (see Table 2). Second, on the basis of
the dependence in Eq. (1), the trailing-edge noise from the blades is
expected to be slightly higher on the upwind side than on the
downwind side (due to the wind speed and rotor tilt angle). This was
conrmed by comparison of the blade noise spectra in states 2a and
2b for the same wind-speed bins (see Sec. III.B.5). Despite the
(relatively) increased nacelle noise, the overall turbine noise is still
dominated by the noise from the blades.
Note that in Fig. 7, the source maximum for the down-going blade
has shifted anticlockwise (relative to the state 2 plots), which can be
explained by the convective amplication factor in Eq. (1) for the
average misalignment angle of 204 deg in this rotor state [25]. Also
note that in the downwind source maps the noise source at the tip of
the up-going blades is more prominent than in the upwind maps. In
the next section it will be shown that, in addition to a small directivity
effect, this difference is mainly due to the lower average wind speed
for these measurements (Table 2), which leads to a lower pitch angle
and higher tip loading (see also Sec. II.A).
B. Noise Sources on the Individual Blades
As mentioned in Sec. I, acoustic eld measurements on the
baseline turbine [25] showed that the average OASPL for the
two standard blades that were used again in the present campaign
(i.e., the baseline and serrated blade) differed by less than 0.05 dB.
Furthermore, it was argued in Sec. II.E that the average level
differences between the blades, as measured with the microphone
array, are accurate within 0.1 dB. Thus, with the present test setup it is
well possible to assess the acoustic performance of the serrations and
the new airfoil. In this section, the possibilities and limitations of
acoustic measurements with a single microphone are rst briey
discussed. Then the array results are used to study the average blade
noise characteristics and the dependence on rotor azimuth, wind
speed, observer position (upwind versus downwind), and rotor state
(tripped versus clean blades).
1. Single-Microphone Analysis
In this section, the acoustic results of a single microphone
are analyzed to demonstrate the possibilities and limitations of
such measurements and to illustrate subjective onsite observations.
During the eld tests, the three different blades could be clearly
distinguished by the difference in swishing noise produced during
Fig. 8 Average stationary source maps for state 2a (clean rotor,
upwind array position). The range of the color scale is 12 dB and the
maximum is adjusted for each frequency band.
Fig. 9 Average stationary source maps for state 2b (clean rotor,
downwind array position). The range of the color scale is 12 dB and the
maximum is adjusted for each frequency band.
-5
-4
-3
-2
-1
0
0 60 120 180 240 300 360
rotor azimuth angl e
∆ψ
∆ψ
(°)
OASPL (dBA)
Baseline
Modi fie d
baseline
blad
e
SIROCCO
blad
e
serrated
blad
e
Fig. 10 Average sound pressure level on central array microphone as a
function of rotor azimuth, for the baseline rotor (2005) and the rotor with
modied blades (state 2a).
1476 OERLEMANS ET AL.
NLR-TP-2009-401
9
the passage of each blade. This can be illustrated by plotting the
OASPL measured on a single microphone (i.e., the reference
microphone at the center of the array) as a function of rotor azimuth
angle (Fig. 10). The overall levels in this gure were summed
between 250 and 800 Hz to focus on the low-frequency noise of the
down-going blade. The result is A-weighted. Furthermore, the levels
were averaged over all measurements and all revolutions in state 2a
(synchronization was done using the trigger signal from the turbine).
As a reference, the result for the test campaign on the baseline turbine
(2005) is shown as well. Note that the results from the baseline rotor
cannot be compared directly with those from the modied rotor, due
to the different meteorological conditions. However, the variation in
noise level between the three blades can be compared for the two test
campaigns. For the baseline rotor, three humps are clearly observed,
representing the swishing noise that is observed when the three
blades pass the 3 oclock position (see also Fig. 8). The three blades
are practically equally noisy, and the amplitude variation (or swish)
during the passage of the blades is about 2.5 dB. For the modied
rotor, three humps with different amplitudes are observed, which can
be associated with the baseline blade, the SIROCCO blade, and
the serrated blade, respectively. Thus, it can be estimated that the
SIROCCO blade yields a reduction of more than 1 dB, whereas the
serrated blade is about 4 dB quieter than the baseline blade. However,
it should be noted that these values only pertain to the low-frequency
noise radiated from around the 3 oclock position, whereas the blades
may also produce signicant noise during the other part of the
revolution, as will be seen subsequently. Moreover, at each moment,
the single microphone picks up the noise from all three blades, and so
the contribution of each blade cannot be extracted from the single-
microphone results. Thus, although Fig. 10 conrms subjective
onsite observations, a dedicated array processing method (as
described in Sec. II.D) is required to obtain a clear picture of the noise
radiated by each individual blade during the complete revolution.
2. Average Blade Noise Characteristics
The source maps for the individual blades, averaged over one
complete revolution and over all measurements in state 2a, are shown
in Fig. 11. The source maps run from 15 to 60 m in radial direction
and have a chordwise extent of 30 m. The black line indicates the
outer 32 m of the blade (trailing edge on upper side). The range of the
scale is 12 dB and the maximum is adjusted for each frequency band,
so that the colors within one row can be compared directly. First, the
source maps show that the differences in source position for the
three blades are small. The source radial position, dened as the
radius at which the maximum level in the source map occurs, is
shown in Fig. 12 as a function of frequency for the three blades.
Because the mesh size of the scan grid was 1 m (Sec. II.D), these
source radial positions are multiples of 1 m. For all blades, the source
radial position increases with frequency, which can be understood
using the relation St f=U const. for the trailing-edge
noise peak [6,33]; for increasing radius, the local blade inow
velocity Uincreases and the boundary-layer displacement thickness
decreases, so that the produced frequencies are higher [24,25].
Except for the lowest frequencies, for which the modied blades
have a lower source radius, the differences in average source radial
positions between the different blades are small. More important, the
source maps in Fig. 11 show that for low frequencies, both modied
blades are signicantly quieter than the baseline blade, especially the
serrated blade. For high frequencies, however, both modied blades
are noisier than the baseline blade, especially the SIROCCO blade.
These trends are illustrated in Fig. 13, which shows the average
integrated spectra for the three blades. These spectra were obtained
by averaging the integrated spectra for all measurements in state 2a,
with equal weights per wind-speed bin. As mentioned in Sec. II.D,
these sound levels represent the contribution of the different blades to
the overall sound pressure level of the turbine, as measured at the
array position. The integrated spectra conrm the low-frequency
noise reduction and high-frequency noise increase for the modied
blades. For the serrated blade, the A-weighted sound pressure level at
high frequencies is even higher than at low frequencies. The reasons
Fig. 11 Average rotating source maps for individual blades in state 2a,
as a function of frequency. The range of the color scale is 12 dB and the
maximum is the same within each row.
32
36
40
44
48
160
315
630
1250
2500
5000
Frequency (Hz)
radius_max (m)
Baseline
Sirocco
Serration
Fig. 12 Average source radial position as a function of frequency for
the three blades in state 2a.
OERLEMANS ET AL. 1477
NLR-TP-2009-401
10
for the increased noise level of the modied blades at high
frequencies will be discussed in subsequent sections. Based on these
average spectra, for the upwind measurements on the clean rotor (i.e.,
state 2a), which are considered most representative for normal
operation and which cover all relevant wind speeds, average overall
noise reductions of 0.5 and 3.2 dB were obtained for the SIROCCO
and serrated blades, respectively. For the other two rotor states,
which covered only the lower wind speeds, average noise reductions
of 0.0 and 1.6 dB (state 1) and 0.2 and 1.2 dB (state 2b) were found
for the SIROCCO and serrated blades, respectively. The reasons for
the lower noise reduction in these states will be discussed in
subsequent sections. The measurement uncertainty of the afore-
mentioned average noise reductions will be discussed subsequently
in Sec. III.B.4.
3. Dependence of Blade Noise on Rotor Azimuth
To better understand the acoustic behavior of the modied blades,
it is interesting to look at the dependence of the blade noise on the
rotor azimuth angle (Fig. 14). This gure shows the overall source
maps for the different blades for 12 rotor azimuth intervals of 30 deg,
starting at 0 deg (12 oclock). These overall source maps
(averaged over all measurements in state 2a) were obtained by
summing the source maps between 160 Hz and 5 kHz, after applying
A-weighting and a correction for array resolution (which depends on
frequency and blade position) to the levels. Thus, the OASPL of the
blade is equal to the sum of all scan levels in a given source map.
Similar to the previous blade source maps, the black line indicates the
outer 32 m of the blade. The range of the color scale is 12 dB, and the
maximum is the same for all maps, so that the colors can be compared
directly. This gure shows that during the downward movement of
the blades, both modied blades are substantially quieter than the
baseline blade. However, during the upward movement, both
modied blades are noisier. These observations are illustrated in
Fig. 15, which shows the integrated blade spectra for the down- and
upward halves of the revolution separately. The high-frequency
noise increase for the modied blades occurs mainly during the
upward part of the revolution, and for the serrated blade, the
increased noise of the up-going blade even dominates the overall
average spectrum.
4. Dependence of Blade Noise on Wind Speed
In addition to the rotor azimuth, the blade noise characteristics
were also found to depend strongly on the wind speed. This is
illustrated in Figs. 16 and 17, which show the integrated blade noise
spectra for the 7 and 10 m=sbins of state 2a, respectively (the state 2b
spectra in Fig. 16 will be discussed subsequently in Sec. III.B.5). The
7m=sblade spectra show a high-frequency hump around 1250
1600 Hz, which is absent in the 10 m=sspectra. As seen in the
previous section, this high-frequency noise is mainly produced
during the upward movement and originates from the very tip of the
blade (Figs. 1114). Because the average blade pitch angle was 0 deg
in the 7m=sbin and higher in the 10 m=sbin, this suggests that the
high-frequency noise at low wind speeds can be associated with the
increased tip loading as a result of the lower pitch angle (see Sec. II.A
for turbine operation details). This also explains the fact that the up-
going blades were noisier in states 1 and 2b than in state 2a (Figs. 7
9), because the average blade pitch angle was lower than in state 2a
(Table 2). Apparently, this tip noise does not follow the trailing-edge
noise directivity function described in Eq. (1), because it is mainly
radiated during the upward movement of the blades. Figure 16 also
shows that the tip noise is much more prominent for the modied
blades than for the baseline blade, which is always dominated by
trailing-edge noise. Thus, the adapted pressure distribution on the
modied blades, possibly in combination with the slightly increased
70
75
80
85
90
95
160
315
630
1250
2500
5000
Frequency (Hz)
SPL (dBA)
Baseline
Sirocco
Serration
5 dB
Fig. 13 Average blade noise spectra for state 2a.
Fig. 14 Average overall source maps for individual blades in state 2a,
as a function of rotor azimuth. The range of the color scale is 12 dB and
the maximum is the same for all maps.
1478 OERLEMANS ET AL.
NLR-TP-2009-401
11
blade loading (Sec. I), changes the tip vortex characteristics such that
tip noise increases.
For the baseline and SIROCCO blades, the frequency of the low-
frequency trailing-edge noise peak is higher at 10 m=sthan at 7m=s,
which can be explained by the lower blade loading (due to the higher
pitch angle), which leads to a thinner suction-side boundary layer.
This is conrmed in Fig. 18, which shows the source radial positions
for both blades in these two wind-speed bins: for a given radius, the
trailing-edge boundary-layer thickness decreases for the higher wind
speed, and so the produced frequencies are higher. For both bins, the
source radial position of the SIROCCO blade is slightly lower than
for the baseline blade, which is consistent with the higher trailing-
edge noise-peak frequency and which suggests that the main
objective of the airfoil design [i.e., to obtain a thinner suction-side
boundary layer (see Sec. I)] has succeeded. If we suppress the
spectral inuence of tip noise by considering only the downward part
of the revolution (Fig. 15), a slight trailing-edge noise increase is
observed between 1 and 3 kHz for the SIROCCO blade, which,
similar to the wind-tunnel results, can be attributed to the increased
pressure-side boundary-layer thickness [35]. However, even if we
consider only the downward part of the revolution, the average
reduction in OASPL for the SIROCCO blade in state 2a is only
1.0 dB, which is lower than the 23 dB found in the wind-tunnel tests.
The reasons for this discrepancy between wind-tunnel and eld
results are not fully clear yet. Apart from blade quality, a possible
explanation could be that instationary inow conditions in the eld
lead to lift uctuations well beyond the prescribed design lift range
[34].
In terms of A-weighted overall sound pressure levels (summed
between 160 Hz and 5 kHz), both modied blades were found to
reach maximum noise levels at a wind speed U10 of about 7m=s,
where the tip loading and therefore the tip noise are highest (the blade
pitch angle only starts to increase signicantly for wind speeds
higher than 7m=s). The noise from the baseline blade, which is
dominated by trailing-edge noise, also peaks around 7m=s. The
corresponding noise reductions (Fig. 19) are lowest around 7m=s
and increase for higher wind speeds, for both the serrated and
SIROCCO blades. Thus, the results indicate that at high wind speeds,
the noise from the three blades is dominated by trailing-edge noise,
which is effectively reduced by the new airfoil shape and the
serrations. However, at lower wind speeds (increased tip loading due
to lower pitch angle), signicant high-frequency tip noise is
generated by both modied blades during their upward movement,
which partly cancels the trailing-edge noise reduction. As a result, the
average noise reductions obtained in states 1 and 2b, which had lower
wind speeds and therefore lower average blade pitch angles
(Table 2), were lower than the 3.2 and 0.5 dB found in state 2a for the
serrated and SIROCCO blades, respectively.
To evaluate these average noise reductions, it is important to
understand the uncertainty levels associated with the measurement
method employed during this study. As argued in Sec. II.E, for the
weather conditions and turbine operation parameters in state 2a, the
measurement uncertainty of the average noise reductions is smaller
72
77
82
87
92
97
160
315
630
1250
2500
5000
Frequency (Hz)
SPL (dBA)
Baseline down
Sirocco down
Serration down
Baseline up
Sirocco up
Serration up
5 dB
Fig. 15 Average blade noise spectra for the upward and downward
part of the revolution in state 2a.
75
80
85
90
95
160
315
630
1250
2500
5000
Frequency (Hz)
SPL (dBA)
Baseline s2a
Sirocco s2a
Serration s2a
Baseline s2b
Sirocco s2b
Serration s2b
5 dB
Fig. 16 Average blade noise spectra for the 7 m=swind-speed bin of
state 2a and state 2b.
70
75
80
85
90
160
315
630
1250
2500
5000
Frequency (Hz)
SPL (dBA)
Baseline
Sirocco
Serration
5 dB
Fig. 17 Average blade noise spectra for the 10 m=swind-speed bin of
state 2a.
32
36
40
44
48
160
315
630
1250
2500
5000
Frequency (Hz)
radius_max (m)
Baseline_U7
Sirocco_U7
Baseline_U10
Sirocco_U10
Fig. 18 Source radial position as a function of frequency for the
baseline and SIROCCO blades in the 7 and 10 m=swind-speed bins of
state 2a.
-1
0
1
2
3
4
5
6
567891011
U10 (m/s)
OASPL (dBA)
Sirocco
Serration
Fig. 19 Overall blade noise reduction as a function of wind speed for
state 2a. The solid lines are third-order least-squares ts through the
measured data.
OERLEMANS ET AL. 1479
NLR-TP-2009-401
12
than 0.1 dB. For different turbine and meteorological conditions
(within the tested range), the uncertainty in the noise reduction can be
assessed from the scatter in Fig. 19: for the serrated and SIROCCO
blades, standard deviations of 0.6 and 0.4 dB were found, which
leads to standard deviations of the mean "for the average noise
reductions of 0.10 and 0.06 dB, respectively. Because these "values
are smaller than the average noise reductions of 3.2 and 0.5 dB, the
average reductions are signicant for the tested range of turbine and
meteorological conditions.
5. Upwind Versus Downwind Measurements
To assess the effect of observer location (i.e., upwind versus
downwind array position), Fig. 16 compares the blade noise spectra
for the 7m=sbin in states 2a and 2b. The meteorological and turbine
parameters were similar for both cases. It can be seen that the level
of the low-frequency trailing-edge noise hump is lower on the
downwind side for all three blades. A partial explanation for this
could be that the downwind integrated blade noise levels suffer from
increased coherence loss (see also Sec. II.E), because the blade noise
propagates through the rotor wake. Indeed, the difference between
the integrated rotor noise level and the level of the reference
microphone was about 1 dB higher for the downwind measurements
than for the upwind measurements in the 7m=sbin. In addition, the
reduced downwind trailing-edge noise levels may be explained by
the dependence in Eq. (1), due to the wind speed and rotor tilt angle
(see also Sec. III.A). In contrast to the low-frequency noise, the high-
frequency tip noise peak has approximately the same level for the up-
and downwind measurements, which means that the relative
importance of tip noise is higher on the downwind side. In addition to
the tip loading effect discussed in the previous section, this increased
importance of tip noise on the downwind side is an additional
explanation for the lower noise reductions obtained in states 1 and 2b
(as compared with state 2a) and for the fact that the up-going blades
are noisier in states 1 and 2b than in state 2a (Figs. 79).
6. Baseline Blade Noise for Different Rotor States
Figure 20 shows the baseline blade noise spectra for the three
different rotor states. With regard to the clean rotor, a number of
differences can be observed between states 2a and 2b: First, the low-
frequency trailing-edge noise peak for state 2a has a higher frequency
and higher level than state 2b. This can be explained, respectively, by
the higher wind speed (i.e., lower blade loading and thinner boundary
layer) in state 2a and the difference in directivity for the up- and
downwind observer positions (see Secs. III.B.4 and III.B.5). Second,
state 2b spectrum shows a high-frequency tip noise peak, which is
almost absent in state 2a. As mentioned before, this can be explained
by the lower average wind speed (and thus higher tip loading) in
state 2b (Sec. III.B.4) and by the increased importance of tip noise on
the downwind side (Sec. III.B.5). Thus, to assess the inuence of
tripping on blade noise, state 1 should be compared with state 2b. The
average conditions for these two states are practically the same,
except for the misalignment angle (Table 2). Figure 20 shows that the
low-frequency trailing-edge noise peak for state 1 has a higher level
and lower frequency than in state 2b. This suggests that the trip has
increased the trailing-edge boundary-layer thickness. Furthermore,
the high-frequency tip noise peak is slightly lower for the tripped
case. Thus, the results do indicate a small effect of tripping (0.6 dB
increase in OASPL for the present conditions), but for a conclusive
answer, measurements should be done with a clean blade and a
tripped blade on one rotor.
IV. Conclusions
Acoustic eld measurements were carried out on a 94-m-diam
wind turbine, with one standard blade, one blade with an optimized
airfoil shape, and one standard blade with trailing-edge serrations.
The blade modications had no adverse effect on their aerodynamic
performance. To assess the effect of blade roughness due to dirt or
insects, the blades were tested in both clean and tripped conditions. A
large horizontal microphone array, positioned at a distance of about
one rotor diameter from the turbine, was used to locate and quantify
the noise sources in the rotor plane and on the individual blades.
Because the array position was xed and the wind direction varied,
both up- and downwind measurements were performed.
The acoustic source maps for the baseline blade showed that for
an observer at the array position, the dominant source was trailing-
edge noise from the outer 25% of the blade. Because of convective
amplication and directivity, practically all noise was produced
during the downward movement of the blade, which caused the
typical swishing noise during the passage of the blades. Both
modied blades showed a signicant trailing-edge noise reduction
at low frequencies, which was more prominent for the serrated
blade. However, the modied blades also showed a noise increase at
high frequencies, which can be associated with the blade tips. This
high-frequency tip noise was mainly radiated during the upward
part of the revolution and was most important at low wind speeds
(i.e., high tip loading) and for the downwind array position.
Nevertheless, for the upwind measurements on the clean rotor,
which were considered to be most representative for normal
operation and covered all relevant wind speeds, average overall
noise reductions of 0.5 and 3.2 dB were obtained for the optimized
blade and the serrated blade, respectively. For both blades, the noise
reduction increased with increasing wind speed. The downwind
measurements on the clean and tripped rotors only covered the
lower wind speeds and showed less noise reduction. Comparison of
the noise from the baseline blade for clean and tripped conditions
suggested a noise increase of 0.6 dB due to tripping.
Acknowledgments
Financial support for this research was given in part by the
European Commissions Fifth Framework Programme, project
reference SIROCCO (Silent Rotors by Acoustic Optimisation)
(ENK5-CT-2002-00702). Financial support was also given by
the Netherlands Organisation for Energy and the Environment
(NOVEM). The authors would like to thank the colleagues from the
University of Stuttgart and from the Netherlands Energy Research
Foundation (ECN) for their valuable contributions to the denition of
the tests and the interpretation of the results.
References
[1] Wagner, S., Bareiss, R., and Guidati, G., Wind Turbine Noise,
SpringerVerlag, New York, 1996.
[2] Guidati, G., Ostertag, J., and Wagner, S., Prediction and Reduction of
Wind Turbine Noise: An Overview of Research Activities in Europe,
AIAA Paper 2000-0042, 2000.
[3] Amiet, R., Acoustic Radiation from an Airfoil in a Turbulent Stream,
Journal of Sound and Vibration, Vol. 41, No. 4, 1975, pp. 407420.
doi:10.1016/S0022-460X(75)80105-2
[4] Howe, M. S., A Review of the Theory of Trailing Edge Noise,
Journal of Sound and Vibration, Vol. 61, No. 3, 1978, pp. 437465.
doi:10.1016/0022-460X(78)90391-7
[5] Blake, W. K., Mechanics of Flow-Induced Sound and Vibration,
Academic Press, New York, 1986.
[6] Brooks, T. F., Pope, D. S., and Marcolini, M. A., Airfoil Self-Noise
and Prediction,NASA Ref. Publ. 1218, 1989.
70
75
80
85
90
160
315
630
1250
2500
5000
Frequency (Hz)
SPL (dBA)
State 1
State 2a
State 2b
5 dB
Fig. 20 Average baseline blade noise spectra for different rotor states.
1480 OERLEMANS ET AL.
NLR-TP-2009-401
13
[7] Dassen, T., Parchen, R., Guidati, G., Wagner, S., Kang, S., and Khodak,
A. E., Comparison of Measured and Predicted Airfoil Self-Noise with
Application to Wind Turbine Noise Reduction,Proceedings of the
European Wind Energy Conference, Irish Wind Energy Association,
Slane, Ireland, Oct. 1997, pp. 422428.
[8] Guidati, G., Bareiss, R., Wagner, S., Dassen, T., and Parchen, R.,
Simulation and Measurement of Inow-Turbulence Noise on
Airfoils,AIAA Paper 97-1698, 1997.
[9] Hutcheson, F. V., and Brooks, T. F., Effects of Angle of Attack and
Velocity on Trailing Edge Noise,AIAA Paper 2004-1031, 2004.
[10] Oerlemans, S., and Migliore, P., Aeroacoustic Wind Tunnel Tests of
Wind Turbine Airfoils,AIAA Paper 2004-3042, 2004.
[11] Moreau, S., and Roger, M., Competing Broadband Noise Mechanisms
in Low Speed Axial Fans,AIAA Paper 2004-3039, 2004.
[12] Herr, M., and Dobrzynski, W., Experimental Investigations in Low
Noise Trailing Edge Design,AIAA Paper 2004-2804, 2004.
[13] Moriarty, P. J., Guidati, G., and Migliore, P., Prediction of Turbulent
Inow and Trailing-Edge Noise for Wind Turbines,AIAA Paper 2005-
2881, 2005.
[14] Grosveld, F. W., Prediction of Broadband Noise from Large
Horizontal Axis Wind Turbine Generators,AIAA Paper 84-2357,
1984.
[15] Glegg, S. A. L., Baxter, S. M., and Glendinning, A. G., The Prediction
of Broadband Noise from Wind Turbines,Journal of Sound and
Vibration, Vol. 118, No. 2, 1987, pp. 217239.
doi:10.1016/0022-460X(87)90522-0
[16] Hubbard, H. H., and Shepherd, K. P., Aeroacoustics of Large Wind
Turbines,Journal of the Acoustical Society of America, Vol. 89, No. 6,
1991, pp. 24952508.
doi:10.1121/1.401021
[17] Lowson, M. V., Theory and Experiment for Wind Turbine Noise,
AIAA Paper 94-0119, 1994.
[18] Fuglsang, P., and Madsen, H. A., Implementation and Verication of
an Aeroacoustic Noise Prediction Model for Wind Turbines,Risø
National Lab., Risø-R-867(EN), Roskilde, Denmark, 1996.
[19] Lowson, M. V., Lowson, J. V., and Bullmore, A. J., Wind Turbine
Noise: Analysis of Results from a New Measurement Technique,
AIAA Paper 98-0037, 1998.
[20] Moriarty, P., and Migliore, P., Semi-Empirical Aeroacoustic Noise
Prediction Code for Wind Turbines,National Renewable Energy Lab.,
Rept. TP-500-34478, Golden, CO, 2003.
[21] de Bruijn, A., Stam, W. J., and de Wolf, W. B., Determination of the
Acoustic Source Power Levels of Wind Turbines,Proceedings of the
European Wind Energy Conference, H. S. Stephens and Associates,
Bedford, England, U.K., Oct. 1984, pp. 889894.
[22] van der Borg, N. J. C. M., and Vink, P. W., Acoustic Noise Production
of Wind Turbines in Practice,Proceedings of the European Wind
Energy Conference, Hellenic Wind Energy Association and European
Wind Energy Association, Oct. 1994, pp. 148155.
[23] Hagg, F., van Kuik, G. A. M., Parchen, R., and van der Borg, N. J. C. M.,
Noise Reduction on a 1 MW Size Wind Turbine with a serrated
Trailing Edge,Proceedings of the European Wind Energy
Conference, Irish Wind Energy Association, Slane, Ireland,
Oct. 1997, pp. 165168.
[24] Oerlemans, S., Sijtsma, P., and Méndez López, B., Location and
Quantication of Noise Sources on a Wind Turbine,Journal of Sound
and Vibration, Vol. 299, Nos. 45, 2007, pp. 869883.
doi:10.1016/j.jsv.2006.07.032
[25] Oerlemans, S., and Schepers, J. G., Prediction of Wind Turbine Noise
and Comparison with Experiment,Proceedings of the Second
International Meeting on Wind Turbine Noise [CD-ROM], Inst. of
Noise Control Engineering, Merseyside, England, U.K., Sept. 2007.
[26] Brooks, T. F., and Burley, C. L., Rotor Broadband Noise Prediction
with Comparison with Model Data,AIAA Paper 2001-2210, 2001.
[27] Schlinker, R. H., and Amiet, R. K., Helicopter Rotor Trailing Edge
Noise,NASA CR-3470, 1981.
[28] Dowling, A. P., and Ffowcs Williams, J. E., Sound and Sources of
Sound, Ellis Horwood, London, 1983.
[29] Howe, M. S., Noise Produced by a Sawtooth Trailing Edge,Journal
of the Acoustical Society of America, Vol. 1, No. 1, 1991, pp. 482487.
doi:10.1121/1.401273
[30] Dassen, T., Parchen, R., Bruggeman, J., and Hagg, F., Results of a
Wind Tunnel Study on the Reduction of Airfoil Self-Noise by the
Application of Serrated Blade Trailing Edges,Proceedings of the
European Wind Energy Conference, H. S. Stephens and Associates,
Bedford, England, U.K., 1996, pp. 800803.
[31] Braun, K. A., van der Borg, N., Dassen, A., Doorenspleet, F., Gordner,
A., Ocker, J., and Parchen, R., Serrated Trailing Edge Noise
(STENO),Proceedings of the European Wind Energy Conference,
James and James Science, London, Oct. 1999, pp. 180183.
[32] Oerlemans, S., Schepers, J. G., Guidati, G., and Wagner, S.,
Experimental Demonstration of Wind Turbine Noise Reduction
Through Optimized Airfoil Shape and Trailing-Edge Serrations,
Proceedings of the European Wind Energy Conference, WIP-
Renewable Energies, Munich, July 2001, pp 351354.
[33] Herr, M., Design Criteria for Low-Noise Trailing Edges,AIAA
Paper 2007-3470, 2007.
[34] Schepers, J. G., Curvers, A., Oerlemans, S., Braun, K., Lutz, T., Herrig,
A., Wuerz, W., Matesanz, A., Garcillán, L., Fisher, M., Koegler, K., and
Maeder, T., Sirocco: Silent Rotors by Acoustic Optimisation,
Proceedings of the Second International Meeting on Wind Turbine
Noise [CD-ROM], Inst. of Noise Control Engineering, Merseyside,
England, U.K., Sept. 2007.
[35] Lutz, T., Herrig, A., Würz, W., Kamruzzaman, M., and Krämer, E.,
Design and Wind Tunnel Verication of Low Noise Airfoils for Wind
Turbines,AIAA Journal, Vol. 45, No. 4, 2007, pp. 779792.
doi:10.2514/1.27658
[36] Herrig, A., Würz, W., Lutz, T., and Krämer, E.: Trailing-Edge Noise
Measurements Using a Hot-Wire Based Coherent Particle Velocity
Method,Proceedings of the 24th AIAA Applied Aerodynamics
Conference, San Francisco, AIAA Paper 2006-3876, June 2006.
[37] Oerlemans, S., Broersma, L., and Sijtsma, P., Quantication of
Airframe Noise Using Microphone Arrays in Open and Closed Wind
Tunnels,International Journal of Aeroacoustics, Vol. 6, No. 4, 2007,
pp. 309333.
doi:10.1260/147547207783359440
[38] Holthusen, H., and Smit, H., A New Data-Acquisition System for
Microphone Array Measurements in Wind Tunnels,AIAA
Paper 2001-2169, 2001.
[39] Wind Turbine Generator SystemsAcoustic Noise Measurement
Techniques, International Electrotechnical Commission Norm 61400-
11, Geneva, 2002.
[40] Johnson, D. H., and Dudgeon, D. E., Array Signal Processing,
PrenticeHall, Upper Saddle River, NJ, 1993.
[41] Sijtsma, P., Oerlemans, S., and Holthusen, H., Location of Rotating
Sources by Phased Array Measurements,AIAA Paper 2001-2167,
2001.
[42] Sijtsma, P., and Stoker, R. W., Determination of Absolute
Contributions of Aircraft Noise Components Using Fly-Over Array
Measurements,AIAA Paper 2004-2958, 2004.
[43] Broersma, L., Acoustic Array MeasurementsAirframe Noise and
Wind Turbine Noise,M.S. Thesis, University of Twente, Enschede,
The Netherlands, Jan. 2006.
J. Wei
Associate Editor
OERLEMANS ET AL. 1481
NLR-TP-2009-401
14
... Serrated or slotted trailing edges have been found to be effective in reducing the dominant trailing edge noise. Oerlemans et al. (2009) modified the trailing-edge serrations and successfully halved the wind turbine noise without adversely affecting the aerodynamic performance. Arce León et al. (2016) and Avallone et al. (2016) both investigated the statistical properties of the boundary-layer flow of a NACA 0018 airfoil with trailing edge serrations by time-resolved stereoscopic particle image velocimetry. ...
... These small vortices vanish without being convected far downstream, and complicated small-scale vortices are observed for the flapped airfoil. These features indicate that the serrated flap can break down the largescale vortex into small vortices, which is believed to be beneficial for noise control (Oerlemans et al., 2009;Moreau and Doolan, 2013;Arce León et al., 2016). ...
Article
Full-text available
Improving the aerodynamic performance of the airfoil is important for optimising the rotor efficiency of the vertical axis wind turbines. As a simple passive control method, the Gurney flap is widely used to improve the aerodynamic performance of airfoils. In this paper, we study the impact of applying a novel serrated gurney flap with different heights on the NACA 0018 airfoil. An improved delayed detached eddy simulation method is adopted to investigate the lift-enhancing mechanism of the serrated gurney flap and the evolution of the downstream vortex system. The results show that the serrated gurney flap can effectively increase the airfoil lift coefficient and the lift-to-drag ratio. The improvement of the serrated gurney flap on the aerodynamic performance of the airfoil is more pronounced at moderate angles of attack. Further analysis of the downstream wake shows that a pair of vortices wraps over both sides of the airfoil and rotates perpendicular to the wake flow, which is produced by the columnar vortex upstream of the flap. These vortices mixed with the wake and accelerated the dissipation of the separated vortex on the suction surface of the airfoil.
... The trailing-edge noise prediction is dependent on serration amplitude (2h) and spanwise wavelength of the serration(λ). Nevertheless, several researchers [33,49,[51][52][53][54][55][56] pointed-out that the model does not give accurate solutions when compared with the measurements, it over-projects maximum noise reduction and does not show increase of the noise beyond cross-over frequency [55]. On the other hand, for more noise reduction to be achieved, another comprehensive study proposed that serration height has to be greater than boundary layer thickness at the trailing-edge (2h ≳ δ) where δ is the boundary layer thickness [57]. ...
... Active control methods act on altering the flow structure such as unsteady pressure fluctuations upstream of the trailing-edge, and passive methods attempt to improve the scattering condition by changing the physical and geometrical properties of the trailing-edge [5,59]. Recently, varies passive control techniques such as the use of porous [59][60][61][62], brushes [63,64], serration [41,48,56], surface treatment [27,65], shape optimization [52], morphing [66] and flexible materials [67] were developed and investigated for the purpose of improving aerodynamic performance as well as reducing noise generated at the trailingedge [62]. In this study, the noise mitigation approach is based on trailing-edge serration, combed trailing-edge and comb-serrated trailing-edge. ...
Article
Global concern about high noise levels in areas near airports and wind farms has generated interest from various groups due to factors such as potential health problems and dissatisfaction among the local community. To accommodate this worthwhile plan of further reducing overall noise levels, some researchers are working on lowering the contribution of trailing-edge noise. The original scientific contribution of this study lies on understanding the efficiency of various trailing edge designs such as baseline, serrations, comb and comb-serrated, across different angles of attack and Reynolds numbers, while also addressing the limitations of existing geometrical models for trailing edges. The study intends to examine the performance of these different configurations, with an emphasis on their effect on acoustic responses. By utilizing large-eddy simulation and applying the Ffowcs-Williams and Hawkings models for noise prediction, an investigation was conducted to assess the impact of these trailing edge configurations on radiated noise at a low Reynolds number of 1.6× 105. The numerical predictions of lift coefficient and surface pressure fluctuations are compared and validated with a published study and experimental data, showing satisfactory results. Further analysis of these study has demonstrated that prominent peaks at lower frequencies (<103) are observed, which are identified as the characteristic frequencies. Moreover, results showed irregular broadband noise (300 - 600 Hz) with increased noise and shifting peak frequency as angle of attack rose. The serrated trailing edge design notably reduced noise levels by roughly 21 dB, especially for low frequencies. Comb-serration increased high-frequency noise by about 9 dB for angles of attack at 0, -1, and -20, and achieved a reduction of approximately 9 dB for angles of attack at 1 and 20. On the other hand, the directivity pattern showed that the maximum noise level is observed to predominantly radiate at an azimuth angle of around 90 degrees for all the cases, ranging from 900 to 2700, indicating that the majority of the source's acoustic energy is being emitted on the suction and pressure sides of the airfoil.
... The problematic of localizing rotating sources was first addressed by Sijstma et al. [14], who proposed to use a rotational speed signal associated with a formulation based on convected moving monopoles. This methodology was successfully applied to helicopter [15], turbofan [16] and wind turbine [17] noise studies. Application to small-scale UAVs configurations present some challenges, indeed their small size and low Reynolds number range require an accurate spatial resolution, especially in the low frequency range associated to BPF and its harmonics. ...
Conference Paper
The radiation of a small-scale low Reynolds number two-bladed rotor of 0, 2 diameter with NACA0012 blade profile in interaction with cylindrical beams at different locations is measured in anechoic room by an array of 45 microphones. Source localization algorithms are used to localize tonal noise sources associated to blade passing frequency (BPF) harmonics that are increased by the presence of the beam in order to better understand noise source generation mechanisms. A first set of BPF harmonics, consisting of a hump centered around 5 * , appears to be localized above the rotor near the blade tip, whereas a second set of BPF harmonics, consisting of a hump centered around 20 − 25 * , seems to have its origin just below the beam in most cases and at ∼ 85% of the blade radius. The increase of rotational speed, beam diameter and the decrease of rotor-beam distance leads to an increase of the first BPF harmonics hump sound levels with similar source locations, except for the noisier rotor-beam configuration for which both humps are located near the blade tip. As this was not consistent with numerical simulations performed on the same configuration which showed that first BPF harmonics levels were associated to unsteady loading on the beam, new simulations were performed to obtain wave fronts between the rotor-beam setup and the array. This allows to explain source locations on maps and to identify which sources were associated to the beam. I. Nomenclature = blade passing frequency () = computational aeroacoustics = conventional beamforming − = CLEAN based on spatial source coherence = cross-spectrum matrix diagonal removal = deconvolution approach for the mapping of acoustic sources = fast Fourier transform − = Ffowcs-Williams and Hawkings = implicit large eddy simulations = nearfield focusing = azimuth angle (•) ′ = point spread functions = rotation per minute (−1) = sound pressure level () = sound power level () = latitude angle (•) = unmanned air vehicles = computational fluid dynamics * Associate professor, ISAE-SUPAERO, Toulouse, France, helene.parisot-dupuis@isae-supaero.fr, AIAA Member. †Associate professor, ISAE-SUPAERO,
... For axial fans, studies by Zenger et al. [18] and Biedermann et al. [19] have demonstrated that introducing serrations to the leading edge can substantially lower fan noise. Similarly, applications in larger-scale settings, such as wind turbines, have been explored by Braun et al. [20] and Oerlemans et al. [21], where serrations reduced overall noise while slightly increasing high-frequency sounds. The potential of serrations extends into the specifics of centrifugal fan design. ...
Conference Paper
Full-text available
This study employs a combined approach of numerical simulation and experimental methods to investigate the flow and noise characteristics of miniature centrifugal fans used in portable electronic devices under free inflow conditions. It compares the effects of serrated structures located at the fan inlet and the blade tips on the flow field and aerodynamic noise. The study utilizes Large Eddy Simulation to capture the flow field within the fan accurately and measures the far-field noise spectrum and directivity. By integrating the pressure fluctuations on the fan blades and upper casing walls in the frequency domain, the research reveals the distribution of dipole sound sources and uses the Q-criterion to visualize the vortex structures inside the fan. The results uncover the flow and noise characteristics of the miniature centrifugal fan and analyze the impact of the serrated structures on these properties. It was found that tonal noise is primarily caused by concentrated intake at the inlet of such miniature centrifugal fans. The contribution of the vortex tongue to noise generation is minimal and predominantly contributes to low-frequency broadband noise. The comparative analysis shows that properly designed serrated structures at the inlet can effectively suppress the generation of BPF tonal noise and that appropriately designed serrated structures at the blade tops can effectively weaken broadband noise and alter the distribution of broadband pressure fluctuations on particular fan casing walls.
... Although it is possible to simulate full wind turbines with PowerFLOW using the actuator line method [23] or by directly simulating the blades [24], simulations including trailing-edge noise of a full-scale turbine would be prohibitively expensive in an industrial environment. Simply performing airfoil section simulations is also not sufficient, as it has been shown [25] that the noise reduction for an airfoil section in a wind tunnel does not correspond to the noise reduction seen in field tests of wind turbines. Hence, we use an approach that uses simulations of airfoil sections, but that predicts the noise these section would produce on a real wind turbine [2,24,26]. ...
Conference Paper
Full-text available
Numerical simulations of a wind turbine blade with and without trailing-edge serrations are validated with full-scale field test of a 130 m diameter onshore wind turbine. Simulations focus on trailing-edge noise and are conducted on extruded airfoil sections of the blade using the lattice-Boltzmann method and very large eddy simulations, which are then propagated to the far-field using the Ffowcs Williams-Hawkings approach, simulating the rotation of the sections and the noise of the entire rotor. Far-field noise spectra at two mean wind speeds are used for validation, with the sound power level of the simulations being within 2.5 dB of field test and the total noise reductions attributed to the serrations being captured within 0.6 dB.
... The results showed that two pairs of counter-rotating vortices were generated near the sawtooth structure, and their strong interaction played a key role in reducing noise. Oerlemans et al. 12 modified the blade of a fan and conducted an experimental study. The findings revealed that the serrated trailing-edge structure significantly reduced the noise less than 1250 Hz by 3-8 dB. ...
Article
Inspired by the silent gliding feather of owl wings, the trailing edge of the duct of a pump-jet propulsor was designed with a similar serrated structure in order to reduce noise generation. Two distinct serrated structures were proposed and evaluated using the detached eddy simulation method with the shear stress transport k−ω turbulence model. The findings indicated that while the hydrodynamic efficiency changed within 1% upon the inclusion of the serrated trailing edge, a significant alteration existed in vortex structures of the wake. More horseshoe and secondary vortices were generated since large-scale vortices induced by the duct were disrupted circumferentially. This phenomenon expedited the distortion and mixing of trailing-edge vortices, causing flow instability. Furthermore, the serrated trailing-edge structure led to noise reduction. Particularly in the 0–1000 Hz range, the sound pressure level behind the duct showed a maximum reduction of 4.43 dB.
... While studies have demonstrated no health-related effects from turbine noise (e.g., Knopper and Ollson, 2011;Radun et al., 2022), residents living nearby often report disturbances from noise, visual impacts, and flickering shadows (Havas and Colling, 2011). In recent years, technological advances in wind turbine design have reduced noise levels (Oerlemans, 2009;Bertagnolio et al., 2023), while noise immissions could be reduced by the establishment of extended distances between the turbines and residential buildings. The presence of a forest, which acts as noise attenuation barrier and as a landscape shield, can also lead to a reduction in the visual effects. ...
Article
Full-text available
Wind power is one of the fastest growing renewable energy sectors and plays a focal role in the transition to a fossil fuel free society in Europe. Technological developments have enabled the construction of turbines within forested areas, which has raised concerns regarding the audiovisual impact on these landscapes. However, there is a paucity of research with regard to the role that forests may play in mitigating the negative impacts of wind farms. In this study, we created a simplified model for noise attenuation based on the ISO 9613-2 and Nord2000 noise models and a visibility model which both relates the audiovisual effect to forest stand structure and applied them in the GIS environment. Our findings suggest that forests can act as effective noise barriers, with the sound attenuation level dependent on the distance that sound travels through the forest, as well as the size and density of the trees. However, in the case of a high elevation sound source (such as wind turbines), the forest begins to act as a noise shield from a distance of between 500 and 1500 m, depending on the height of the forest and the land topography. While current noise models do not consider the impact of tree species, our visibility model accounts for tree size, density and species, as well as understorey and thinning. Our results indicate that spruce trees provide a better visual constraint whereas visibility distances within mature Calluna-type pine forests tend to be more extensive. Both models include variables that can be adjusted by forest management, thereby allowing integration with forest planning software. Overall, this study presents indicative methods for the evaluation of potential forest landscape shields, a concept that could have broad applications, including Landscape Value Trading.
Article
An experimental campaign to study the impact of a distinct type of vortex generator — rod type (RVG), on the flow characteristics and the acoustic far-field pressure of a wind turbine airfoil, is conducted. Airfoils exhibit decreased aerodynamic performance at high inflow angles due to turbulent boundary layer flow separation. RVGs are applied to mitigate the flow separation. However, this benefit is accompanied by an acoustic penalty. An assessment of the impact of RVGs on the far-field noise emission is conducted for the DU96-W-180 airfoil. The evolution of the boundary layer impacted by the rods is analyzed through Particle Image Velocimetry (PIV) measurements. The resulting reduction in the separation zone is observed through oil flow visualization. Analysis of the sound spectrum for airfoils with/without RVGs is conducted for a range of frequencies (300 Hz to 4000 Hz). Results show a reduction of the noise level at relatively low frequencies, at the expense of an increased noise level in the mid-high frequency ranges. While the former is caused by the reduction of the flow separation, the latter is determined by the combined contribution of the noise scattered by the RVG and by the change in boundary layer characteristics at the airfoil trailing edge.
Article
This study employs a comprehensive combination of experimental and numerical methodologies to delve into the aeroacoustic attributes of a small horizontal axis wind turbine with optimized blades. The experimental investigation is conducted within a semi-anechoic chamber, where both original and optimized geometry models are meticulously positioned to measure the sound pressure levels across a range of rotational speeds and positions. In parallel, the numerical simulations employed the large eddy simulation, complemented by the Ffowcs Williams–Hawkings analogy, facilitating detailed examinations of both aerodynamic and acoustic aspects in the original and optimized modes. The findings reveal a subtle enhancement in aerodynamic performance with the optimized serrated blade configuration when compared to the original. Nevertheless, the reduction in noise levels within the frequency domain was remarkable, culminating in an impressive overall sound pressure reduction of about 10 dB. Furthermore, an intriguing observation emerged from noise measurement in acoustic room: the noise production experiences a marked escalation as the turbine rotational speed intensifies, particularly within the downstream domain. The lateral noise level is found to be lower compared to the axial direction and the reduced noise emission for the serrated optimized blade is more dispersed in the plane of rotation than the original blade, which was pointed out to be nearly uniform. The results provide valuable insights into the interplay of aerodynamics and aeroacoustics in the context of small wind turbines with optimized blades.
Article
Trailing edge serration is an effective method to reduce broadband noise generation by an airfoil. However, the noise reduction performance can be significantly reduced when there is flow misalignment at the serration. This experimental study investigates the impact of a shift in the serration position downstream of the airfoil trailing edge on the noise reduction performance, referred to in this paper as serration extension. Experiments were performed on a 100-mm-chord NACA 0012 wing model with sawtooth trailing edge serrations. The serration performance was studied at 0° and 7° flap-down configurations at various angles of attack. The serrations with three different extension lengths of 5, 10, and 15 mm were tested and compared with the baseline case without extension. The emitted noise was measured with a phased microphone array. The results show a significant reduction in broadband high-frequency noise by extension under loading conditions. Particle image velocimetry measurements of root flow along the wall-normal plane reveal diminished crossflow across the serration after extension. Furthermore, the extensions cause an up-to-25% increase in the maximum lift coefficient and improved or unaltered lift-to-drag ratios.
Article
Within a parametric study on brush-type trailing-edge extensions, the noise reduction potential of several design concepts was determined. The obtained database represents the first phase of an ongoing project with the longterm objective to develop scaling laws for a future application of such devices as add-on solutions for today's aircraft components. The experiments comprised both acoustic and aerodynamic measurements on a zero-lift generic plate model (Re = 2.1 x 10_6 to 7.9 x 10_6) in DLR's open jet Aeroacoustic Wind Tunnel Braunschweig. Noise data were taken by means of a directional microphone system. Measurement results indicate a significant source reduction potential in excess of 10dB, depending on the configuration. Two relevant noise reduction mechanisms were identified: 1) the suppression of narrowband bluntness noise, as well as 2) the reduction of broadband turbulent boundary-layer trailing-edge noise.
Conference Paper
This paper reports new analysis and prediction development of rotor broadband noise. The two primary components of this noise are blade-wake interaction (BWI) noise, due to the blades' interaction with the turbulent wakes of the preceding blades, and "Self" noise, due to the development and shedding of turbulence within the blades' boundary layers. Emphasized in this report is new modeling and code development for rotor blade self noise. The analysis and validation employs data from the HART program, a model BO-105 rotor test conducted in the German-Dutch Wind Tunnel (DNW). The BWI noise predictions are based on measured pressure response coherence functions using cross-spectral methods. The self noise predictions are based on previously reported semi-empirical modeling obtained from isolated airfoil sections but include new modeling and code developments for higher speeds, inclusion of all Doppler effects and capability to define performance and local blade segment flow conditions and positions from advanced comprehensive rotorcraft analysis such as CAMRAD.Mod1. Both BWI and self noise from individual blade segments are now Doppler shifted and summed at the observer positions. Prediction comparisons with measurements show good agreement for a range of rotor operating conditions from climb to steep descent. The broadband noise predictions, along with those of harmonic and impulsive Blade-Vortex Interaction (BVI) noise predictions, demonstrate a significant advance in predictive capability for main rotor noise.
Conference Paper
The paper tackles the problem of aerodynamic noise from wind turbines. It concentrates on the specific noise sources which are relevant for a large, modem wind turbine in the 1 MW range. This is mainly trailing-edge noise which originates from the interaction of the turbulent boundary layer around the blade with the trailing edge. Additional noise sources are blade tip noise and inflow-turbulence noise while low-frequency noise is considered to be less relevant. The physical mechanisms of the noise generation are explained and prediction models for the single sources are described. These models are believed to be sensitive enough to capture the effect of airfoil geometry correctly. Therefore, they might be used to consider aerodynamic noise already during the design process, i.e. to design less noisy wind turbines. A further topic is the reduction of aerodynamic noise, mainly the serrated trailing edge which is the most promising concept to reduce trailing-edge noise and consequentlyt he total noise of a wind turbine. In the last part of the paper a new approach to the prediction of aerodynamic noise in general is presented. Here some relevant concepts of aeroacoustic theory are explained.
Conference Paper
Aeroacoustics wind tunnel tests are performed for six airfoils that are candidates for use on small wind turbines. A microphone array was used to locate and quantify leading and trailing edge noise from the airfoils. The broadband sound levels were found to scale with U 4.5 and the inflow turbulence noise levels were found to scale with U 6. With regard to the test set-up, it was found that a treatment of porous material at the model-endplate junctions yielded a broadband extraneous noise reduction of up to 10 dB.
Conference Paper
This paper describes a hot-wire based method for two-dimensional trailing edge noise measurements mainly for use in wind tunnels with high background noise levels. The method is based on the cross correlation of two hot-wire signals, allowing to measure the Coherent Particle Velocity (CPV) of the emitted sound waves. The strong directional sensitivity of the hot-wires leads to a suppression of parasitic noise, which significantly improves the signal-to-noise ratio. Due to the small dimensions of the sensors together with the assured laminar flow condition they are well suited for inflow measurements. To obtain quantitative results in terms of sound pressure level, the sensitivity of the measurement setup is derived by simulation of the response of the hot-wires to a line source. For validation, trailing edge noise measurements were performed at 60 m/s and a Reynolds number of 1.6 × 106 in the closed test section of the Laminar Wind Tunnel Stuttgart (LWT), using the CPV-method. Finally, the same airfoils were investigated in the open jet of the Aeroacoustic Wind Tunnel Braunschweig (AWB) using a phased microphone array. The quantitative comparison of the experimental results obtained in the two wind tunnels required the application of appropriate wind tunnel corrections. The obtained sound pressure frequency spectra are basically found to be parallel in the frequency range of sufficient measurement accuracy. The total sound pressure levels vs. lift coefficient show a more or less constant offset of about 2 dB between AWB and LWT. Given the totally different measurement principles this can be regarded as a good agreement. Finally results of a NACA 0012 airfoil are presented and compared to published data.