Vous êtes sur la page 1sur 8

Testing the Performance of a

Ground-based Wind LiDAR


System
One Year Intercomparison at the Offshore
Platform FINO1
B. Caadillas, A. Westerhellweg, T. Neumann; DEWI GmbH, Wilhelmshaven

B. Canadillas

English

Introduction
Traditional meteorological masts that are commonly used in
wind energy applications are limited and very expensive to
install in offshore locations. Therefore, it is necessary to
search for an alternative method to replace the standard
wind measurements (cup/sonic anemometers, vanes) used
on a mast. The LiDAR (Light Detection And Ranging) technique has become a reasonable alternative in the last years
with the advantage that higher vertical resolution measurements over the whole rotor diameter of a wind turbine can
be performed [1], [2]. However, before taking lidar systems
as an accepted wind measurement alternative, we have to
validate as well as understand how this technique computes
both mean and turbulent wind parameters.
One important issue is the time resolution at which the wind
measurements have to be sampled which will depend mostly on the application for which the wind measurements will
be used. For instance, for wind resource and power curve
assessments, a 10-min interval is recommended by the
standards [3], [4]; however several investigations pointed
out that a higher resolution (1 Hz or even less) is needed for
a better definition of the power curve [5]. In addition, monitoring the rapid changes in the wind conditions is important
for optimal operation of wind farms.
58

DEWI MAGAZIN NO. 38, FEBRUARY 2011

In the last years, several field experiment intercomparisons


with in-situ sensors located on meteorological masts have
proven the lidar system to be a reliable instrument for
obtaining mean values (10-min). In particular, similar lidar
systems as the one used in this study (Leosphere WindCube)
have been tested over homogeneous terrain (e.g., [6], [7])
with excellent results. However such comparisons have only
relied on 10-min averaged data.
Within the joint research project RAVE (Research at
AlphaVEntus), the RAVE-LIDAR work package deals with
remote sensing measurements using lidar systems at the
German Offshore test site [8]. Part of the work carried out by
DEWI is presented in this article. This study aims to examine
the performance of a wind lidar measurement located at
FINO1 site, using the mast to provide reference measurements.
The present study is split into two major parts. In the first
part, an intercomparison between lidar and conventional
mast-based instruments, such as cup/sonic and vanes, is
performed based on 10-min data. Several wind parameters
are considered, namely wind speed, wind direction and turbulence intensity. In the second part, we complement the
10-min data intercomparison with a spectral analysis in
order to get a more complete picture of the flow in the
ranges between one hour and about one second.

The Field Measurement Campaign


A ground-based pulsed lidar system, the so-called
WindCube, developed and manufactured by the French
company Leosphere has been used in this study. The lidar
system was positioned on a container roof at approximately 10 m distance to the northwest of the offshore research
mast FINO1 (Fig. 1) and performs continuous measurements since July 2009. The FINO1 platform (lat. 540.86N,
long. 635.26E), equipped with a 100 m lattice tower, is
operating since September 2003 in the German Bight
approx. 45 km to the northwest of the island of Borkum. In
this study, the FINO1 mast is used as ground truth to provide a reference to which the lidar data shall be compared.
The mast instrumentation used here (Tab. 1) are cup anemometers at 101.5 m (top-mounted) and at 90 m, 80 m
and 70 m located on the South-East side of the mast, ultrasonic sonic anemometer at 80 m and wind vanes at 90 m
and 70 m located on the northwest side of the mast.
Slow-response profile instrumentation is sampled once
per second (1 Hz) and stored as 10 min averages. Fastresponse instrumentation is sampled with a frequency of
10 Hz. Moreover, east of FINO1 the Alpha Ventus offshore
wind farm is located at a distance of 405 m to the nearest
wind turbine. The farm consists of 6 REpower 5M wind
turbines with a hub height of about 92 m LAT and a rotor
diameter of 126 m and 6 Areva Wind M5000 wind turbines
with a hub height of about 91 m LAT and a rotor diameter
of 116 m. The first wind turbine was installed on 2009-0715 and went into operation on 2009-08-12. However, for
the time period used in this study not all wind turbines
were in operation at the same time, so that wind coming
from the east may be disturbed by those wind turbines for
some periods.
The WindCube Lidar System
A Doppler wind lidar used for atmospheric purposes, measures the frequency shift (Doppler shift) of narrow bandwidth laser light backscattered from microscopic particles
or aerosols in the air being transported by the wind. The
frequency shift is proportional to the velocity of the scattered target along the propagation path of the laser beam
or Lines-Of-Sight (LOS) which is estimated with a specific
signal processing algorithm implemented in the device. The
range to the target is determined by the time of traveling
back and forward.
The WindCube Doppler lidar system used in this study
emits 200 ns pulses of 10 J energy at a wave length of 1.54
m [9]. There is a direct interface to the FINO-network,
which allows the data transfer, as well as the control of the
device by using remote control software. The measurements were set-up at heights from 70 to 101 m every 10m,
from 100 to 200 m every 20 m and at 250 m. The lowest
height is 70 m because the height of the lidar device is
already 23 m msl and the Windcube has a minimum measurement height of 40 m.
From lidar measurements, the estimation of the wind
velocity vector (u, v, w) requires retrieval of the LOS velocity (also called as radial velocity vr). The Windcube, located
at the FINO1 platform, uses the Velocity- Azimuth-Display
(VAD) technique of lidar scanning (conical scan lidar beam
at a fixed elevation angle) to derive the 3D components

Power Quality
DEWI carries out measurements
and evaluations to determine the
electrical characteristics of single
wind turbines and of wind farms
according to the currently applicable
standards (e.g. IEC 61400-21) and is
accredited by German Accreditation
Council in line with EN ISO/IEC
17025:2005 and by MEASNET.
As one of the leading international
consultants in the field of wind energy,
DEWI offers all kinds of wind energy related
measurement services, energy analyses and
studies, further education, technological,
economical and political consultancy for
industry, wind farm developers, banks,
governments and public administrations.
DEWI is accredited to EN ISO/IEC 17025 and
MEASNET for certain measurements and is
recognised as an independent institution in
various measurement and expertise fields.

DEWI MAGAZIN NO. 38, FEBRUARY 2011

www.dewi.de

59


Fig. 1:

Tab. 1:

Lidar position at the offshore research platform FINO1

Sensor Type
cup, vane (Vector A100 LK, Thies Wind Vane Classis)
cup, sonic (Vector A100 LK, Solent 1210R3-50)
cup, vane (Vector A100 LK, Thies Wind Vane Classis)
Cup (Vector A100 LK)

Overview of the FINO1 measurements involved in the intercomparison

w
Vr180

Vr0

Height msl [m]


70 m
80 m
90 m
101.5 m

Vr90
v
u

Vr270
Y

Fig. 2:

Sketch of the scan method performed by the


WindCube lidar

Fig. 4

Wind speed ratio lidar/cup (curves shifted).


Sector between black vertical lines is the sector
used for the intercomparison. Red circular
areas show approx. the mast shadow on the
lidar measurements

60

DEWI MAGAZIN NO. 38, FEBRUARY 2011

Fig. 3:

Data availability at FINO1 for the period from 2009-08-01 to


2010-07-31

Height msl
[m]

(y = m x)
m [-]

a [-]

101.5

0.998

0.997

90

0.988

0.991

R2

Bias
(vlidar-vcup)
[m/s]

0.010

0.998

-0.030

0.997

(y = a x +b)
b [m/s]

Counts
[-]

Remarks

-0.02

0.13

0.13

6770

dir sector [190;240]

-0.13

0.22

0.18

6831

dir sector [190;240]

80

0.992

0.997

-0.051

0.997

-0.08

0.19

0.17

6928

dir sector [190;240]

80-USA

1.003

1.003

0.007

0.996

0.03

0.19

0.19

6704

dir sector [190;240]

70

0.997

1.000

-0.036

0.997

-0.03

0.16

0.16

6981

dir sector [190;240]

Tab. 2:

Statistics of the filtered horizontal wind speed.

Fig. 5:

Scatter plot of horizontal wind speed of lidar against sonic


measurements at 80 m

Height [m]

Fig. 6:

a [-]

b []

R2

Bias
(dirlidar-dircup)
[]

(y = a x +b)

Scatter plot of lidar wind direction against sonic wind direction


at 80 m

Counts

Remarks

90

1.013

-0.106

0.995

2.7

2.87

0.18

6831

dir sector [190;240]

80-USA

1.019

-2.140

0.996

1.9

2.12

0.19

6667

dir sector [190;240]

70

1.010

-1.455

0.995

0.8

1.30

0.16

6928

dir sector [190;240]

Tab. 3:

Statistics of the filtered wind direction

of the wind. This technique is based on the following radial


velocity equation:
vr=ucos()sin() + vsin()sin() + wcos()
where u, v and w are the wind vector components and and
are the azimuthal and zenithal angles of the wind vector.
The WindCube lidar performs four successively radial velocity measurements at azimuth-angle intervals of 90 around
the circle formed by a conical scanning, i.e. at =0, =90,
=180 and =270 and at a fixed elevation angle , as illustrated in Fig. 2.
The limitations of this approach include the assumption of
horizontal homogeneity of the wind field over the sensed
height. Its temporal resolution for acquiring a full 3-D wind
vector is about 4.6 sec. (0.22 Hz), i.e. each revolution takes
about 4.6 seconds. However, the last three radial velocity
values plus a new measured value are used to derive wind
speed and direction profiles.
Experimental Results
Intercomparison Based on 10-min Averaged Values
In this section, an intercomparison study based on 10-min

values of the measured quantities is performed. The variables used to describe and compare the sensors are wind
speed, wind direction and standard deviation of the horizontal wind speed. In order to determine how well both data
groups (lidar and FINO1) are statistically related to each
other, a correlation analysis is carried out. Moreover, the
criteria used to statistically describe the accuracy and precision of the lidar in comparison to a reference sensor are:
Sample bias (B), Comparability (C), Precision or standard
deviation of the differences (S).
Data Availability
The evaluated data cover a one year measurement period
from 01 August, 2009 to 31 July, 2010. The availability of the
different measuring instruments at the different heights is
shown in Fig. 3. For the cup anemometers it is close to 100
% at all heights. The sonic anemometer shows a lower
availability with approx. 97.5 %. For the lidar two different
curves are displayed. The overall availability of the lidar is
given by the blue line, i.e. here, all 10-min mean values were
included in which at least one single value exists. The black
graph shows the availability considering only 10-min data
DEWI MAGAZIN NO. 38, FEBRUARY 2011

61

Fig. 7:

Scatter plot of lidar wind standard deviation against sonic (left) and turbulence intensity (right) against mean wind speed at 80 m

with 100% availability. At heights comparable with the FINO1


data the lidar availability amounts to 98 %. The availability
decreases with increasing height to 91 % at a height of 200
m, respectively 83 % at 250 m. This is due to the decreasing
of the signal to noise ratio with height.
Mast Shadow on Measurements
In Fig. 4 the ratio of wind speed measured by the lidar and
cup as a function of the wind direction at 90 m is depicted for
all compared heights. The curves are shifted vertically
artificially for better visualization. As can be seen, the mast
shadow effects on the anemometers and laser beams are
clearly noticeable at certain wind directions. The plot shows
mast shadow effects on the lidar beams at approximately
40, 110, 180 and 270. These are approximately the
directions where one of the four lidar beams is on mast
shadow. These directions are slightly shifted for different
sensing heights.
Moreover the graphs show the lightning cage (90, 180,
270 and 360) for the height at 101.5m and the upwind and
downwind flow retardation (280-340) for heights at 90 m,
80 m and 70 m. The shadow effect from the neighbouring
wind turbines from Alpha Ventus wind farm is also
appreciable.
Data Filtering
In order to ensure the data comparability between the lidar
and reference instruments, the following filtering criteria
are applied before the analysis:
Rain: Events with rain periods are discarded.
Lidar availability: Only lidar data with an availability of
100% for each 10 min value are used.
Signal to noise (CNR) threshold: Values with CNR < -22 dB
are discarded.
Wind speed: Only mean wind speeds in the range of 4-16
ms-1 are considered in the comparison. This corresponds
to the range where the cup/sonic anemometers were
calibrated.
Wind direction: For the wind speed comparison and for
the heights of the boom mounted anemometers only

62

DEWI MAGAZIN NO. 38, FEBRUARY 2011

wind directions within the range interval 190-240 are


used which avoids the wakes from the wind turbines, and
the direct mast shade on the anemometers and lidar
beams (see Fig. 4).
After filtering the 10-min dataset is reduced to 6981 values
(13%, direction sector 190-240, 70 m msl).
Wind Speed
Values of correlation coefficient, slope, offset, bias, comparability, precision and number of observations for all heights
are tabulated in Tab. 2. At each of the compared heights a
scatter plot of 10-min averaged horizontal wind speed is
performed, however, as an example, only the one at 80 m is
displayed in Fig. 5 (left). Two versions of the linear regression
analysis (with (y=ax+b) and without (y=mx) offset) are considered. Correlation analysis suggests that the wind speeds
from the lidar system agree well with the tower measurements with values of R higher than 0.99 for all heights. The
estimates of the biases show that at almost all heights
(except for the sonic) the lidar wind speed tends to be
smaller than the tower wind speed (negative bias). Values of
C and S are similar and fall between 0.19 and 0.13 ms-1.
Considering the different sampling and averaging strategies
of the two measurement systems, very good agreement can
be noticed.
Wind Direction
Tab. 3 summarizes the results of the wind direction comparison. Biases for all heights are positive, which could be
attributed to an inaccuracy in the north alignment during the
lidar setup. The comparison of the wind direction measurements shows a good correlation between lidar and vanes
with a coefficient of correlation higher than R = 0.99. A
Scatter plot at 80 m, using the sonic anemometer to measure the wind speed direction, is depicted in Fig. 6.
Wind Standard Deviation
Before the computation of the turbulence intensity (TI=U/U),
comparisons of the 10-min averages of the standard deviation of the horizontal wind speed (U) at all investigated

Run-mean
Run-mean
filter
filter

Fig. 8:

Comparison (sonic (blue) versus lidar (red)) of horizontal wind speed (left) and radial velocity (right) spectra at 80m. In this example (right), the
sonic data were projected onto a vector aligned into the 90 beam direction

heights are carried out. A Scatter plot for the sonic at 80 m is


depicted in Fig. 7 (left). Unlike the wind speed comparison,
the scatter of the standard deviation is larger.
The WindCube Lidar system measures higher values of U
than the cup anemometers at all observing heights (positive
bias). In contrast, in [11] for instance, it was shown that using
another lidar system (ZephIR) smaller values of U (approximately 20%) are retrieved as a consequence of the spatial
and temporal averages applied to the lidar measurements.
The exact nature of this discrepancy found in this intercomparison has not yet been determined. Some possible reasons
for the discrepancy include data processing technique.
The curves represented in Fig. 7 (right) show the turbulence
intensity as a function of the wind speed at 80 m height. The
highest values for the turbulence intensity arise with very
low wind velocities. There is a constant decrease up to the
range of approx. 12 m/s. For higher wind speeds the turbulence intensity increases because the roughness of the sea
increases. For all wind speeds however, the lidar measures
turbulence intensity higher than the sonic anemometer. A
quite similar picture results for all heights investigated.
Spectral Analysis
The 10-min comparison addressed in the previous section
provides a measure of the mean characteristics of the turbulent flow where higher frequencies of the wind are filtered
out, although there is considerable information about the
higher frequency data with relevance to wind energy application including, among others, wind load studies and wind
turbine design.
Here the Power Spectrum Density (PSD) of wind velocity
fluctuations is evaluated. We are trying to understand from
the turbulence point of view how both measurements correlate to each other and to explain why turbulence measured with this LiDAR shows higher values with respect to insitu measurements. For this comparison study measurements from the sonic anemometers at 80 m are used.
To perform the spectral analysis, the Welch algorithm has
been used. For further information, please refer to [12]. The
power spectrum from the lidar is compared to the power

spectrum from the sonic anemometer at 80 m, and the


results are illustrated in Fig. 8 (left). To enhance the readability of the spectra, a log-log scale is used. Note that the spectra represent an average over all observed wind speeds for
the period selected. Atmospheric stability for this period was
mostly near-neutral. Winds were coming from the northwest
sector.
In Fig. 8 (left), the cut-off frequencies differ because of different time resolutions of sonic (10Hz) and lidar measurements
(0.67 Hz). The black dashed line represents the theoretical
slope of a Kolmogorov inertial sub-range (f-2/3). The green
dashed vertical line corresponds to a frequency of 1/600
(10min). The loop at the end of the lidar spectrum (about f>
0.21Hz) is a consequence of the temporal averaging algorithm which acts as a run-mean filter with a time window of
about 4.6 sec. (time which laser beam takes to scan a full
circle). Note that this lidar system cannot predict frequencies
higher than about 0.21 Hz and, therefore, above that value it
is better not to derive any conclusions from the spectrum.
As can be clearly seen, the lidar spectrum presents a signal
increase at frequencies between 0.004 and 0.2 Hz. The cause
of this apparent energy increase is unknown at present, but
could be related to instrument noise, differences in sampling
methods, or algorithm methods to construct the velocity
vector. However, it is shown in the left hand of this figure that
there is a good agreement between both wind spectra at low
frequencies.
It is well known that the shape of the turbulence spectra
depends among others on the thermal stratification, the
height above the ground surface, and the wind velocity. In
[12] an analysis according to different ranges of wind speed,
wind direction or atmospheric stability was performed and it
showed that any of these parameters seem to have a direct
influence on the lidar spectra shape.
One of the averaging mechanisms applied during the
WindCube lidar measurement is the probe length which
leads to an averaging of the wind component along the laser
beam within a sampling width of 20 m. In Fig. 8 (right), we
focus on the radial velocity intercomparison in order to avoid
the algorithm methods used to derivate the horizontal wind

DEWI MAGAZIN NO. 38, FEBRUARY 2011

63

Height [m]

(y = a x +b)

Bias
(lidar-cup)
[m/s]

Counts
[-]

Remarks

a [-]

b [m/s]

R2

101.5

1.069

0.030

0.834

0.06

0.13

0.12

6770

dir sector [190;240]

90

1.043

0.028

0.841

0.05

0.12

0.11

6831

dir sector [190;240]

80

1.034

0.028

0.851

0.05

0.12

0.11

6928

dir sector [190;240]

80-USA

0.973

0.053

0.789

0.04

0.14

0.13

6704

dir sector [190;240]

70

1.024

0.027

0.867

0.04

0.11

0.10

6981

dir sector [190;240]

Tab. 4:

Statistics of the filtered standard deviation of the horizontal wind speed.

speed and thus to check whether the anomalous shape of


the lidar spectra at certain frequencies can be a consequence
of this velocity derivation algorithm or not.
In order to compare the radial velocity along the laser beam,
the 10Hz sonic data are projected onto a vector aligned with
the same azimuth and elevation angle as the lidar beam
(line-of-sight). Fig. 8 (right) suggests that there is a good
agreement between both radial wind spectra which suggests
that the probe-length averaging seems not to have an influence on the shape of the horizontal wind lidar spectrum.
Conclusion
A Lidar measurement has been performed over a period of a
whole year with an availability of 98% for the heights
involved in the intercomparison, proving that lidar can be a
reliable wind measurement device even in harsh offshore
weather conditions. Even for the highest measurement level
of 250m and availability of 83% could be achieved, which
proves that the lidar device is an adequate instrument to
span the whole rotor area of modern offshore wind turbines.
In general good agreements are found between both data
sets when comparing 10 min averaged measurements. The
comparison of the 10 min lidar wind speed data with mast
data shows a high correlation with R2 higher than 0.99 for all
heights.
One important result is that the mean wind deviations
between lidar and cup measurements are smaller than the
uncertainties of the mast measurements.
In contrast to the onshore case, the offshore wind conditions
are governed by small wind shear and small turbulence
intensity, which are ideal conditions for the lidar technique in
VAD mode, which depends on homogeneity in the spatial
and temporal scanning area. Hence the mean wind speed
and wind direction can be measured with high quality; however the regarded lidar type (Leosphere Windcube) overestimates the turbulence intensity at the present stage.
The lidar spectra present a signal increase at certain frequencies, which is independent on wind direction, wind speed or
stability conditions. The cause of this apparent energy
increase is unknown at present, but could be related to
approximations in sampling or algorithm methods to construct the velocity vector. For the global turbulence this
effect leads to a systematic overestimation. On the other
hand, the radial wind spectra comparison suggests that the
probe-length averaging seems not to have an influence on
the shape of the horizontal wind lidar spectrum.

64

DEWI MAGAZIN NO. 38, FEBRUARY 2011

In this study the processing techniques used by the WindCube


lidar to derive the horizontal wind speed have not been analyzed. It is beyond the scope of this study to find out whether
the lidar algorithms are responsible for the spectra behavior
and/or can be optimized to compensate this error.
Acknowledgements
Within the joint research project RAVE (Research at
AlphaVEntus), the RAVE-LIDAR project was set up to prepare
the application of remote sensing measurements using lidar
systems at the German Offshore test site [8]. Part of the work
that is carried out by DEWI is presented in this report. We
also like to thank Leosphere for their support.

References:
[1] Courtney, M., R. Wagner, und P. Lindelw. Commercial lidar profilers
for wind energy. A comparative guide. EWEC. 2008.
[2] Emeis, S., M. Harris, und R. M. Banta. Boundary-layer anemometry by
optical remote sensing for wind energy applications. Meteorologische
Zeitschrift 16, Nr. 4 (2007): 337-347.
[3] IEC61400-1. Wind turbines-part 1: Design requirements. Technical
report, International Electrotechnical Commission (IEC).
[4] IEC61400-12-1. Power performance measurements of electricity producing wind turbines. 2005.
[5] Anahua, E., M. Lange, F. Boettcher, S. Barth, und J. Peinke.
Characterization of the Wind Turbine Power Performance Curve by
Stochastic Modeling. EWEC. Athens, 2006.
[6] Albers, A., and A. Janssen. Windcube evaluation report. Technical
report, Deutsche WindGuard Consulting GmbH, 2008.
[7] Gottschall J., and M.l Courtney. Verification test for three WindCubeTM
WLS7 LiDARs at the Hvsre test site. Technical report, Ris-R-1732(EN),
2010.
[8] Rettenmeier, A., et al. Development of LiDAR measurements for the
German Offshore Test Site. 14th International Symposium for the
Advancement of Boundary Layer Remote Sensing. IOP Conf. Series:
Earth and Environmental Science, 2008.
[9] Pauliac, Romain. WindCube. Users Manual. Leosphere, June 2008.
[10] Mann, J, et al. Comparison of 3D turbulence measurements using three
staring wind lidars and a sonic anemometer. Ris-R-1660(EN), 2008,
135140.
[11] Rozenn Wagner, Torben Mikkelsen, Michael Courtney. Investigation of
turbulence measurements with a continuous wave, conically scanning
LiDAR. EWEC. 2009. 10
[12] Canadillas, B., A. Bgu, T. Neumann. Comparison of turbulence spectra
derived from LiDAR and sonic measurements at the offshore platform
FINO1. DEWEK. 2010. Germany.

200m ultra-portable wind profiler

Werbung
Kunde
1/1
s/w oder 4c
Location 1
Period 1

Location 2
Period 2

Location 3
Period 3

S IT E A S S E S S M E N T > PR E -E VALUATION > MICR O - S I TI N G > POW ER C U R V E V ER I FI C ATI O N

200m

Location 4
Period 4

The original WINDCUBE is reducing uncertainty worldwide with units


on every continent. You asked for the same comprehensive wind
proles, only from a small, lighter, lower-power unit. We listened.
The new WINDCUBE v2 lidar remote sensor offers:
Ultra-portable, easy to deploy, lowest cost of ownership
Bankable data, to reduce your project nancing costs
Complement met mast data with variable measurement heights up

60m

to 200m
Unattended, low-power, quiet operation in any weather or noise

Download Free ROI Papers.

conditioneven off shore

www.lidarwindtechnologies.com/ROI

Satellite communications from anywhere in the world

LidarWindTechnologies, S.A.

Our global partnership provides sales, technical support and customer service worldwide, wherever your wind site location.

Live demonstrations

EWEC

AW E A

CanWEA

Vous aimerez peut-être aussi