You are on page 1of 22

Journal of Wind Engineering

and Industrial Aerodynamics 90 (2002) 1381–1402

Wind models for simulation of power fluctuations


from wind farms
Poul Srensena,*, Anca D. Hansena, Pedro Andre!
Carvalho Rosasa,b
a
Wind Energy Department, Ris National Laboratory, P.O. Box 49, DK-4000 Roskilde, Denmark
b
Department of Electrical Power Engineering, Technical University of Denmark, DK-2800 Lyngby,
Denmark

Abstract

This paper presents a wind model, which has been developed for studies of the dynamic
interaction between wind farms and the power system to which they are connected. The wind
model is based on a power spectral description of the turbulence, which includes the coherence
between wind speeds at different wind turbines in a wind farm, together with the effect of
rotational sampling of the wind turbine blades in the rotors of the individual wind turbines.
Both the spatial variations of the turbulence and the shadows behind the wind turbine towers
are included in the model for rotational sampling. The model is verified using measured wind
speeds and power fluctuations from wind turbines.
r 2002 Elsevier Science Ltd. All rights reserved.

Keywords: Wind models; Turbulence; Wind turbines; Power fluctuations; Tower shadow effect; 3p;
Coherence

1. Introduction

This paper presents a wind model, which has been developed to support studies of
the dynamic interaction between large wind farms and the power system to which
they are connected, and to support improvement of the electric design of wind
turbines as well as grid connections.
The work is mainly a result of a Danish project titled ‘‘Simulation of wind power
plants’’, but also a Brazilian Ph.D. study titled ‘‘Power quality and stability issues of
integration of large wind farms’’ has contributed to the model development. This
Ph.D. study is accomplished in Denmark.

*Corresponding author.

0167-6105/02/$ - see front matter r 2002 Elsevier Science Ltd. All rights reserved.
PII: S 0 1 6 7 - 6 1 0 5 ( 0 2 ) 0 0 2 6 0 - X
1382 P. Srensen et al. / J. Wind Eng. Ind. Aerodyn. 90 (2002) 1381–1402

As part of the Danish project, a dynamic model of a Danish wind farm has been
implemented in the dedicated power system simulation tool DIgSILENT, compris-
ing models of grid, wind turbines and wind speeds. General descriptions of the
applied models have been given by Srensen et al. [1], whereas the present paper will
concentrate on the wind models.
The background for the Danish project and the Brazilian Ph.D. study is the fast
development and ambitious targets for wind energy. As the wind energy
development is concentrated in areas with good wind resources, where the power
grids are often not so strong, wind energy is becoming a significant source of supply
to the power systems in these areas. As a consequence, wind energy is also playing an
increasingly important role in the operation of these power systems.
The increased wind energy development is also reflected in the requirements for
grid connection of wind turbines. National standards for power quality of wind
turbines have been supplemented by a new IEC 61400-21 [2] standard for
measurement and assessment of power quality of grid connected wind turbines.
The wind models presented in this paper have been developed with the intention to
support simulation of the power fluctuations, which are quantified in IEC 61400-21.
Such simulations can help the wind turbine industry to improve the electric design of
wind turbines, and to reduce the costs for grid connection to what is required from
the point of view of keeping a high power quality in the systems.

2. Wind model structure

The wind model is essential to obtain realistic simulations of the power


fluctuations during continuous operation of the wind farm. The present wind model
combines the stochastic effects caused by the turbulence and deterministic effects
caused by the tower shadow.
The stochastic part includes the (park scale) coherence between the turbulence at
different wind turbines as well as the effects of rotational sampling, which is known
to move energy to multiples (often denoted p’s, e.g. 3p) of the rotor speed from the
lower frequencies [3].
Only the longitudinal component of the wind speed is included, which is normally
a reasonable assumption for wind turbines, because this component has the
dominating influence on the aero loads on wind turbines.
The park scale coherence is included, because it ensures realistic fluctuations in the
sum of the power from all wind turbines, which is important for the maximum power
output from the wind farm. In IEC 61400-21, it is specified that the maximum 200 ms
average power as well as the maximum 1 min average power must be measured as a
part of the power quality test of wind turbines.
To be able to extend the results from measurement on a single wind turbine to a
wind farm, IEC 61400-21 assumes that the turbulence of the wind at the different
wind turbines is uncorrelated. Moreover, it is assumed that for each wind turbine i;
the maximum 200 ms power P0:2;i appears at rated power Pn;i : These assumptions
leads to the following estimate of the maximum 200 m s power P0:2S of a wind farm
P. Srensen et al. / J. Wind Eng. Ind. Aerodyn. 90 (2002) 1381–1402 1383

with N wind turbines


rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
X
N XN
P0:2S ¼ Pn;i þ ðP
i¼1 0:2;i
 Pn;i Þ2 : ð1Þ
i¼1

Assuming that the turbulence at different wind turbines is uncorrelated, a


corresponding relation can be made to predict the standard deviation of the power
from a wind farm, knowing the standard deviation of the power from each wind
turbine. However, analyses of measurements, e.g. Tande et al. [4], have shown that
the assumption of uncorrelation implies an underestimation of the standard
deviation of the power from a wind farm.
The main reason for this underestimation is that the actual correlation between the
turbulence at different wind turbines has some influence, particularly when the
distance between the wind turbines is small. The present wind speed model includes
the coherence of the turbulence to be able to obtain better estimates of maximum
power and power standard deviation.
The effect of the rotational sampling is included because it is a very important
source to the fast power fluctuations during continuous operation of the wind
turbine. The fast fluctuations are particularly important to assess the influence of
wind turbines on the flicker levels in the power system, i.e. the level of voltage
fluctuations causing flicker in the illumination from electric light bulbs.
In many cases, e.g. Srensen [5], measurements have shown that the 3p effect
due to rotational sampling provides the main contribution to flicker emission
from wind turbines to the grid during continuous operation of the wind
turbines.
The structure of the wind model is shown in Fig. 1. It is built as a two-step model.
The first step of the wind model is the park scale wind model, which simulates the
wind speeds vhub;1 ; y; vhub;N in a fixed point (hub heigh) at each of the N wind
turbines, taking into account the park scale coherence. The second step of the wind
model is the rotor wind model, which includes the influence of rotational sampling
and integration along the wind turbine blades as the blades rotate. The rotor wind
model provides an equivalent wind speed veq;i for each wind turbine i; i.e. a single
time series for each wind turbine, which is conveniently used as input to a simplified
aerodynamic model of the wind turbine.
The park scale wind model is implemented in an external program PARKWIND
that generates a file with wind speed time series, which are then read by DIgSILENT.
This model interface is possible because the wind speeds are assumed to be
independent on the operation of the wind farm, which is reasonable because the
required wind speed vhub;i for each wind turbine i is the wind speed in hub height if
wind turbine i was not erected.
The present version of PARKWIND does not include the effects of wakes in the
wind farm, but the mean wind speed and turbulence intensity could be modified to
account for these effects. Jensen [6] suggested a wind farm model to predict the
reduction of the mean wind speed in a wake relative to the ambient mean wind speed.
Frandsen and Thgersen [7] suggested a model for combining the ambient
1384 P. Srensen et al. / J. Wind Eng. Ind. Aerodyn. 90 (2002) 1381–1402

vhub,1 Rotor veq,1 Wind


wind turbine 1
θWTR,1
Park scale wind model
vhub,2 Rotor veq,2 Wind
wind turbine 2
θWTR,2

vhub,n Rotor veq,n Wind


wind turbine n
θWTR,N

Fig. 1. Structure of wind model.

turbulence and the wake-induced turbulence to predict the influence of the wind
farm wakes on fatigue loadings on the wind turbines.
The rotor wind models describes the influence of rotational sampling and
integration along the wind turbine blades as the blades rotate. The present model for
the wind field includes turbulence as well as tower shadow effects. The wind shear is
not included in the model, because it only has a small influence on the power
fluctuations as discussed in Section 4.1.
The rotor effects are included for each of the n wind turbines individually. The
wind speed seen by the rotating blades of the ith wind turbine depends on the
azimuth position yWTR;i of the wind turbine rotor. As illustrated in Fig. 1, yWTR;i is
fed back from the mechanical part of the wind turbine model.
The wind model provides an equivalent wind speed veq;i for each wind turbine i;
which is used as input to the aerodynamic model of that wind turbine. veq;i is a single
time series for the wind turbine i; which takes into account the variations in
the whole wind speed field over the rotor disk. The advantage of using the equivalent
wind speed is that it can be used together with a simple, Cp -based aerodynamic
model, and still include the effect of rotational sampling of the blades over the
rotor disk.
P. Srensen et al. / J. Wind Eng. Ind. Aerodyn. 90 (2002) 1381–1402 1385

3. Park scale model

3.1. Review of methods

Different methods can be applied to simulate the wind speeds in a wind farm.
Estanqueiro [8] used the Shinozuka [9] method based on a cross spectral matrix.
Initially, Mann’s [10] simulation method was suggested used in this project, because
the calculation speed of that method is generally faster than cross spectral matrix
method. Some of the results using Mann’s model were presented in Srensen
et al. [11].
Both the Estanqueiro application of the Shinozuka method and the Mann method
assume Taylor’s frozen turbulence hypothesis illustrated for a two-dimensional wind
speed field in Fig. 2. The wind speed field is generated in spatial dimensions in the
first place, and then the field is moved forward with the mean wind speed.
Taylor’s frozen turbulence hypothesis is a reasonable assumption for simulations
where the simulated time series only pass the object once, like a wind turbine rotor.
Both the Shinozuka method (with frozen wake) and the Mann method are used in
computer programs for simulation of mechanical loads on wind turbines to simulate
the wind speed variations in the rotor plane of a single wind turbine. In that case, the
simulated wind speed time series only pass the object once.
But for park scale wind simulations, the Taylor hypothesis is not so realistic,
especially when the wind direction is along a line of wind turbines. In that case,
simulations assuming Taylor’s frozen turbulence will generate wind speed time series
which are identical for all wind turbines in the line, apart from a delay corresponding
to the travel time for the wind from one wind turbine to the other.
One consequence of assuming frozen turbulence is that the coherence between the
turbulence in two points on a line in the wind direction will be one, which is not
realistic. Another measure, which reveals the error introduced by assuming Taylor’s

V0

Fig. 2. Simulation of park scale wind speeds with the assumption of Taylor’s frozen turbulence
hypothesis.
1386 P. Srensen et al. / J. Wind Eng. Ind. Aerodyn. 90 (2002) 1381–1402

frozen turbulence is the cross correlation function. These aspects are discussed
further in Section 5. The assumption of frozen turbulence will significantly affect the
summation of power fluctuations from the wind turbines in a line.

3.2. The complex cross spectral method

To avoid the assumption of Taylor’s frozen turbulence, we have introduced a new


method for simulation of park scale wind speeds. The new method is based on
Shinozukas cross spectral matrix method, and it uses a complex cross spectral matrix
as Shinozuka originally proposed instead of the real matrices that have been used in
frozen turbulence models for wind turbines. The cross spectral matrix becomes
complex as a result of the time delay between points with longitudinal components of
their separation.
The new method also has the advantage that it does not produce more data than
what is needed. The Mann method on the other hand produces a grid of data, which
then in turn has to be interpolated in at the initial positions of the wind turbines
relative to the data grid. The complex cross spectral method directly generates a
single time series at the position of each wind turbine. Because of this data reduction,
the new method also reduces the computation time considerably compared to the
fast Mann method.
The cross spectral method is based on the cross power spectrum matrix Sðf Þ;
which with N points (corresponding to a wind farm with N wind turbines) is an
N  N matrix. We have chosen to use the frequency in Hz, f. Each element Src ðf Þ in
row r; column c of Sðf Þ is determined as the cross power spectrum between the
turbulence at wind turbine number r and c: Src ðf Þ is defined as the Fourier transform
of the cross correlation function Rrc ðtÞ according to
Z N
Src ðf Þ ¼ Rrc ðtÞej2pf t dt: ð2Þ
N

The cross correlation function Rrc ðtÞ between vhub;r ðtÞ and vhub;c ðtÞ is defined as
Rrc ðtÞ ¼ Efvhub;r ðtÞ  vhub;c ðt  tÞg; ð3Þ
where Eff ðtÞg denotes the mean value of f ðtÞ over the time t:
The first step in the cross power spectral method is to determine Src ðf Þ: Fig. 3
shows the two points r and c each corresponding to a wind turbine.
The distance between the two wind turbines is drc ; with the angular direction yrc
from north. The mean wind speed V0 and the wind direction yV are also shown.
arc ¼ yV  yrc is the inflow angle.
Fig. 3 also indicates the delay time trc for the wind field to travel from wind
turbine c to wind turbine r: Simple geometry yields trc determined as
cosðarc Þ  drc
trc ¼ : ð4Þ
V0
To represent the time delay trc in the cross power spectrum, we assume that Rrc ðtÞ
as defined in Eq. (3) is symmetric about trc : Then using Eq. (2), it can be shown that
P. Srensen et al. / J. Wind Eng. Ind. Aerodyn. 90 (2002) 1381–1402 1387

drc

θrc
αrc
θV

r V0
V0·τrc
Fig. 3. Two points r and c each corresponding to a wind turbine. The distance between the two points r
and c is drc ; with a direction yrc from north. V0 is the mean wind speed, and yV is the wind direction. arc is
the resulting inflow angle, and trc is the delay time.

the argument of the complex number Src ðf Þ is 2pf trc ; i.e.


Src ðf Þ ¼ jSrc ðf Þj  ej2pf trc ; ð5Þ
where j is the complex unity number. The size of the cross power spectrum jSrc ðf Þj
can be determined using the standard definition of the coherence function
g2 ðf ; d; V0 Þ;
jSrc ðf Þj2
g2 ðf ; drc ; V0 Þ ¼ : ð6Þ
Srr ðf ÞScc ðf Þ
Combining Eqs. (6) and (5), we can express the complex cross power spectrum as
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Src ðf Þ ¼ gðf ; drc ; V0 Þ Srr ðf ÞScc ðf Þ  ej2pf trc : ð7Þ
The second step is to discretise the frequency to be able to represent the spectra in
a numeric computer code. Simulating time series with the period length TP ; the
frequency f is discretised in steps Df ¼ 1=TP :; i.e. the ith frequency f ½i ¼ iDf . The
corresponding discrete value of Src ðf Þ is Src ½i ¼ Src ðfi Þ  Df :
We also discretise the time by the sampled representation of the wind speeds as
time series with time steps Dt ¼ 1=fs ; where fs is the sampling frequency in Hz. This
sampling limits the frequency to 7fs =2 and consequently the frequency index i to
7Ns =2; where
N s ¼ T P  fs ð8Þ
is the number of samples in the simulated time series. Obviously, Ns must be an
integer, and preferably an exponent of 2 which enables the use of an FFT to speed up
the Fourier transformation used in the end of the method. This can be obtained by
adjusting either TP or fs according to Eq. (8).
Selecting an appropriate sampling frequency and assuming two-sided spectra, this
discretisation ensures that the variance s2r of the wind speed at wind turbine r is
1388 P. Srensen et al. / J. Wind Eng. Ind. Aerodyn. 90 (2002) 1381–1402

preserved according to
Z N
2
sr ¼ Srr ðf Þ df
N
Z fs =2
E Srr ðf Þ df
fs =2
N
X s =2

E Srr ½i : ð9Þ
i¼Ns =2

The discretisation is only done for frequency indices iX0; because the values for
io0 are given by Src ½i ¼ Src ½i ; where denotes complex conjugation.
The third step is for each frequency index iX0 to resolve the discrete matrix S½i
with the elements Src ½i into a product of the transformation matrix H½i and the
transpose of its conjugate H T ½i ; i.e.
S½i ¼ H½i H T ½i : ð10Þ
Choosing the solution where H½i is a lower triangular matrix, i.e. the element
Hrc ½i ¼ 0 if c > r; the elements can be found one by one. The diagonal elements are
determined according to
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Xr1
Hrr ½i ¼ Srr ½i  H ½i Hrk ½i ð11Þ
k¼1 rk

and the elements below the diagonal are determined according to


P
Src ½i  c1
k¼1 Hrk ½i Hck ½i
Hrc ½i ¼ : ð12Þ
Hcc ½i
It can be seen from Eqs. (11) and (12) that H½i gets the same phases as S½i ; i.e.
zero phase shift in the diagonal and a phase shift 2pf trc below the diagonal,
corresponding to the delay of wind speed between two wind turbines r and c.
The fourth step is for each frequency index iX0 to generate an N  1 vector E½i of
unity complex numbers with a random phase. This is done by simulating N random
phase angles jr ½i using a random generator with uniform distribution in the interval
½0; 2p ; and calculate each element Er in row r of E according to
Er ¼ ejjr ½i : ð13Þ
The fifth step is for each frequency index iX0 to calculate a vector V hub ½i
containing the ith Fourier coefficients of all N wind speed time series according to
V hub ½i ¼ H hub ½i E½i : ð14Þ
The imaginary part of V hub ½0 should be set to zero, because an imaginary part of
this frequency component does not make sense.
Finally, for each wind turbine r; the Fourier coefficients are joined in an array
Vhub;r ; and an inverse Fourier transform is performed to obtain the time series
vhub;r ðtÞ:
P. Srensen et al. / J. Wind Eng. Ind. Aerodyn. 90 (2002) 1381–1402 1389

3.3. Spectral distributions

At the present state, the model is capable of simulating wind speeds with power
spectra of either Kaimal or Hjstrup type, but implementation with another spectrum
is straightforward, as the spectra are used explicitly in the model according to Eq. (7).
The Kaimal spectrum has been selected in the first place because it is used widely,
while the Hjstrup spectrum was selected because it includes more energy than the
Kaimal spectrum at the low frequencies, and this has shown to agree better in a
number of cases.
The two-sided Kaimal [12] spectrum SKai ðf Þ is determined according to
52:5ðz=V0 Þ
SKai ðf Þ ¼ u2 ð15Þ
ð1 þ 33ðz=V0 Þf Þ5=3
and the Hjstrup [13] spectrum SHoj ðf Þ according to
!
2:5ðAHoj =V0 Þ 52:5ðz=V0 Þ 1
SHoj ðf Þ ¼ u2  5=3
þ 5=3  :
1 þ 2:2ððAHoj =V0 Þf Þ 1 þ 33ðz=V0 Þf 1 þ 7:4ðz=AHoj Þ
ð16Þ
In Eqs. (15) and (16), we have used the (hub)height z and the friction velocity u
which can be determined by the roughness length z0 according to

z kV0
ln ¼ ; ð17Þ
z0 u
where k ¼ 0:4 is the von Karman constant. The spectra in Eqs. (15) and (16) are
given as they are developed in the lower boundary layer. However, wind turbines are
far above the lower boundary layer today. For large wind turbines, it is often
assumed that the length scale LE20z only increases with height up to z ¼ 30 m, e.g.
Danish code of practice for loads and safety of wind turbine structures [14].

3.4. Coherence

As it is seen from Section 3.2, it is simple to use any coherence function with the
PARKWIND method. We have chosen to implement a Davenport [15] type
coherence, and use the decay factors recommended by Schlez and Infield [16] as
default values in the program.
Schlez and Infield studied the horizontal two-point coherence for separations
greater than the measurement height, and their recommendations are based on
estimates on own measurements and several other measurements.
The Davenport type coherence function between the two points r and c (see Fig. 3)
can be defined in the square root form
gðf ; drc ; V0 Þ ¼ earc ðdrc =V0 Þf ; ð18Þ
where arc is the decay factor. Schlez and Infield uses a decay factor which depends on
the inflow angle arc shown in Fig. 3. The figure shows that arc ¼ 0 corresponds to
1390 P. Srensen et al. / J. Wind Eng. Ind. Aerodyn. 90 (2002) 1381–1402

points separated in the longitudinal direction, and arc ¼ 901 corresponds to points
separated in the lateral direction.
With a given arc ; the decay factor can be expressed according to
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
arc ¼ ðalong cos arc Þ2 þ ðalat sin arc Þ2 ; ð19Þ

where along and alat are the decay factors for separations in the longitudinal and the
lateral directions, respectively. Using our definition of coherence decay factors in
Eqs. (18) and (19), the recommendation of Schlez and Infield can be rewritten as to
use the decay factors
s
along ¼ ð1575Þ  ; ð20Þ
V0

alat ¼ ð17:575Þðm=sÞ1  s; ð21Þ


where s is the standard deviation of the wind speed in m/s.

4. Rotor wind model

4.1. Equivalent wind speed

As described in Section 2, the rotor wind model provides an equivalent wind speed
veq ; which takes into account the variations due to turbulence and tower shadow in
the wind speed field over the rotor disk. This section describes how the equivalent
wind speed is derived from the rotor wind speed field.
Fig. 4 shows the three bladed wind turbine with the wind field vðt; r; yÞ: It is seen
that the positions are given in the polar coordinates ðr; yÞ; where y is denoted the
azimuth angle.
The aerodynamic torque Tae ðtÞ is given as the sum of the blade root moments
Mb ðtÞ in the drive direction of each blade b; i.e.
X
3
Tae ðtÞ ¼ Mb ðtÞ: ð22Þ
b¼1

Fig. 4 indicates that the blades are profiled from the inner radius r0 to the outer
radius R of the rotor disk. Linearising the blade root moment dependence on the
wind speed we obtain
Z R
Mb ðtÞ ¼ MðV0 Þ þ cðrÞðvðt; r; yb Þ  V0 Þ dr; ð23Þ
r0

where MðV0 Þ is the steady state blade root moment corresponding to the mean wind
speed V0 ; and cðrÞ is the influence coefficient of the aero load on the blade root
moment in radius r:
For aero torque, a typical load distribution along the blades can be obtained by
assuming cðrÞ to be proportional to r and r0 ¼ 0:1R; which has also been assumed in
P. Srensen et al. / J. Wind Eng. Ind. Aerodyn. 90 (2002) 1381–1402 1391

v(t,r,θ)

(r,θ)

(r0,θ1)
(R,θ1)
1

Fig. 4. The wind speed field in the rotor plane is given as vðt; r; yÞ; the blades are profiled from the inner
radius r0 to the outer radius R: The figure also indicates azimuth position y1 of blade number 1.

the implemented model. However, for the sake of completeness, we will formally
keep cðrÞ and r0 here.
Inserting Eq. (23) in Eq. (22) we obtain
3 Z R
X
Tae ðtÞ ¼ 3MðV0 Þ þ cðrÞðvðt; r; yb Þ  V0 Þ dr: ð24Þ
b¼1 r0

Now we define the equivalent wind speed veq ðtÞ as the single wind speed time series
which would give the same aerodynamic torque as the actual wind speed field, i.e.
veq ðtÞ must fulfil
3 Z R
X
Tae ðtÞ ¼ 3MðV0 Þ þ cðrÞðveq ðtÞ  V0 Þ dr: ð25Þ
b¼1 r0
1392 P. Srensen et al. / J. Wind Eng. Ind. Aerodyn. 90 (2002) 1381–1402

Combining Eq. (25) with Eq. (24) gives the equivalent wind speed a the mean
value of contributions from all three blades:
1X 3
veq ðtÞ ¼ vc ðt; yb Þ; ð26Þ
3 b¼1

where we have used the weighted wind speed vc ðt; yb Þ defined as


RR
r cðrÞvðt; r; yb Þ dr
vc ðt; yb Þ ¼ 0 R R : ð27Þ
r0 cðrÞ dr

Eqs. (26) and (27) express the equivalent wind speed as a weighting of all the wind
speeds which are instantaneously seen by the wind turbines along the blades.
We could now simulate the wind speeds in a number of points in the rotor plane
like it is typically done in codes for simulation of mechanical loads on wind turbines.
However, a much more computer time saving simulation method for simulation of
the equivalent wind speed has been developed in Ris National Laboratory.
This simulation method was first presented by Langreder [17] for the contribution
from turbulence. Later, Rosas et al. [18] included tower shadow effects in the
method. In this paper, it is combined with the park scale model.
The equivalent wind speed simulation method is based on Riss frequency domain
models [19–21]. These models are based on expansion of the wind speed field in the
rotor plane in the azimuth angle.
To understand the simulation model, we first expand the weighted wind speed for
a single blade in the azimuth angle, i.e.
X
N
vc ðt; yb Þ ¼ v*c;k ðtÞejkyb ; ð28Þ
k¼N

where v*c;k ðtÞ is the kth azimuth expansion coefficient of vðt; yb Þ determined according
to
Z
1 2p
v*c;k ðtÞ ¼ vc ðt; yb Þejny dy: ð29Þ
2p 0
Inserting Eq. (28) in Eq. (26) yields
XN
veq ðtÞ ¼ v*c;3k ðtÞej3kyWTR ; ð30Þ
k¼N

where yWTR ¼ y1 is the wind turbine rotor position obtained from the mechanical
model.
It is seen from Eq. (30) that only the azimuth expansion coefficients with orders
which are multiple of 3 contribute to the sum. This is because of the symmetric
structure of rotor, which causes the contributions from the other orders to even out
when they are summed up for all three blades. If a 1p and/or 2p are still significant in
a measurement of torque or power, this is often an indication that the blades are not
pitched with exactly the same angle.
P. Srensen et al. / J. Wind Eng. Ind. Aerodyn. 90 (2002) 1381–1402 1393

In the present implementation of the rotor wind model, we have only included the
0th and 3rd harmonics, i.e. Eq. (30) has been approximated to
veq ðtÞE v*c;0 ðtÞ
þ 2Ref*vc;3 ðtÞgcosð3yWTR Þ
þ 2Imf*vc;3 ðtÞgsinð3yWTR Þ: ð31Þ
The idea of the rotor wind model is to simulate the azimuth expansion coefficients
v*c;k ðtÞ in the first place as independent on the azimuth position of the rotor, and then
use Eq. (31) to generate the equivalent wind speed which includes the azimuth
dependence.
The azimuth expansion coefficients v*c;k ðtÞ are sums of contributions from the
turbulence model, tower model and the mean wind speed. The mean wind speed
contributes with V0 to v*c;0 ðtÞ: The contributions from the turbulence model and the
tower model are determined in the next subsections.
Other effects like wind shear and yaw error could be included, but these effects
mainly contribute to the 1p; which is filtered away by the summation of the 3
symmetric blades as mentioned above.

4.2. Turbulence model

As a consequence of the description above, the turbulence model generates the


azimuth expansion coefficients v*c;k;turb ðtÞ of the turbulence field. It has been shown
that [19,21] the power spectral density (PSD) Sv*c;k ðf Þ of v*c;k;turb ðtÞ can be determined
according to
Sv*c;k ðf Þ ¼ Fv*c;k ðf Þ  Sv ðf Þ; ð32Þ
where Sv ðf Þ is the PSD of the wind speed in a fixed point, and Fv*c;k ðf Þ is denoted the
admittance function. Fv*c;k ðf Þ can be determined by a triple integral, which can be
resolved into the double integral
RR RR
r r cðr1 Þcðr2 ÞFk ðf ; r1 ; r2 Þ dr1 dr2
Fv*c;k ðf Þ ¼ 0 0 RR ð33Þ
ð r0 cðrÞ drÞ2
and the single integral
Z 2p qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1
Fk ðf ; r1 ; r2 Þ ¼ glv ðf ; r21 þ r22  2r1 r2 cosðyÞÞ cosðnyÞ dy: ð34Þ
2p 0

Here, glv ðf ; dÞ is the square root coherence function between two points with a
distance d in the rotor plane. glv ðf ; dÞ is assumed to be the same horizontally (i.e.
laterally) and vertically in the plane.
Eqs. (33) and (34) have been solved numerically by Srensen [21]. Using the
Laplace operator s ¼ jo ¼ j2pf ; Langreder [17] defined the transfer functions
Hv*c;k ðj2pf Þ with the size defined as
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
jHc;k ðj2pf Þj ¼ Fv*c;k ðf Þ ð35Þ
1394 P. Srensen et al. / J. Wind Eng. Ind. Aerodyn. 90 (2002) 1381–1402

and used the numerical results to fit Hv*c;0 ðj2pf Þ and Hv*c;3 ðj2pf Þ to linear filters.
Defining the constant d ¼ R=V0 ; the results of these fittings are
0:99 þ 4:79ds
Hc;0 ðsÞE ; ð36Þ
1 þ 7:35ds þ 7:68ðdsÞ2

0:0307 þ 0:277ds
Hc;3 ðsÞE : ð37Þ
1 þ 1:77ds þ 0:369ðdsÞ2
Using vhub ðtÞ from the park scale model as input and a linear filter with the transfer
function Hc;0 ðj2pf Þ; v*c;0;turb ðtÞ is now simulated according to
V* c;0;turb ðf Þ ¼ Hc;0 ðj2pf Þ  Vhub ðf Þ; ð38Þ
where V* c;0 ðf Þ and Vhub ðf Þ are the Fourier transforms of v*c;0;turb ðtÞ and vhub ðtÞ;
respectively.
v*c;3;turb ðtÞ is a complex variable, and it was shown [19,21] that the real and
imaginary parts are uncorrelated with each other and with azimuth expansion
coefficients of other orders k: Distributing the variance between the real and
imaginary parts of v*c;3;turb ðtÞ evenly, they are determined by the relations between
Fourier transforms
1
RefV* c;3;turb ðf Þg ¼ pffiffiffi Hc;3 ðj2pf Þ  V3;Re ðf Þ; ð39Þ
2

1
ImfV* c;3;turb ðf Þg ¼ pffiffiffi Hc;3 ðj2pf Þ  V3;Im ðf Þ; ð40Þ
2
where V3;Re ðf Þ and V3;Im ðf Þ are Fourier transforms of uncorrelated stochastic signals
with the same PSD as the wind speed in a fixed point.
To support the simulation of V3;Re ðf Þ and V3;Im ðf Þ; Langreder also fitted a filter
which converts uniformly distributed white noise to a signal with the PSD as the
Kaimal spectrum.

4.3. Tower shadow model

Today most wind turbines are constructed with a rotor upwind of the tower to
reduce the tower interference of the wind flow. Early wind turbines often had lattice
towers, but because of the visual impact, tubular towers are the most common today.
The tubular towers have more effect on the flow than lattice towers. In the upwind
rotor case, the tower disturbance vtow can be modelled using potential flow theory.
Ekelund [22] found
x 2  y2
vtow ¼ V0 a2 ; ð41Þ
ðx2 þ y2 Þ2
where a is the tower radius, and x and y are the components of the distance from
each blade to the tower centre in the lateral and the longitudinal directions,
respectively.
P. Srensen et al. / J. Wind Eng. Ind. Aerodyn. 90 (2002) 1381–1402 1395

Rosas [18] used Eqs. (41), (27) and (29) to calculate the azimuth expansion
coefficients caused by the tower shadow. Neglecting the effect of the blade bending,
these coefficients become constants v*c;k;tow ; which can be added to contributions
from the turbulence.

5. Verification

The parkscale model has been verified using wind speed measurements on two sea
masts SMW and SMS on the Vindeby offshore site. The offshore site is shown in
Fig. 5.
Eleven wind turbines are installed on this site, and three meteorological masts
were installed as part of a major data collection on the worlds first offshore
wind farm. The distance between the two sea masts SMS and SMW was
dSW ¼ 807 m, and the angle ySW ¼ 2971: The measurements used in this paper were
acquired in 1994.
Fig. 6 shows 1 h of 1 min block average values of wind speed measurements with
the mean wind speed V0 ¼ 11 m/s and the mean wind direction yV ¼ 2931; together
with PARKWIND simulations with the same mean wind speed and wind direction.

1500

1000

500

SMW
0
Wind turbines

SMS Met. masts


-500

-1000

-1500
LM

-2000
-500 0 500 1000
Fig. 5. Vindeby offshore site with 11 wind turbines, two offshore masts SMW and SMS and the land mast
LM.
1396 P. Srensen et al. / J. Wind Eng. Ind. Aerodyn. 90 (2002) 1381–1402

Measured
14
SMW
wind speed [m/s]

SMS
12

10

8
0 10 20 30 40 50 60
time [min]

Simulated
14
SMW
wind speed [m/s]

SMS
12

10

8
0 10 20 30 40 50 60
time [min]
Fig. 6. Simulation of wind speeds compared to measurements on two sea masts in Vindeby offshore wind
farm.

Both measurements and PARKWIND indicate a delay of SMS with a little more
than 1 min relative to SMW. This corresponds very well to the expected delay for
11 m/s with 807 m distance.
A more consistent verification is to compare the coherence functions. However,
this requires a substantial data set, because the coherence in 807 m distance is very
small.
Fig. 7 shows the square root coherence of 6  6.5 h wind speed with average wind
direction 2911. To get an idea of the variation of the wind direction, the standard
deviation of the 234 10-min mean values of wind directions is 4.31, with maximum
wind direction 3001 and minimum 2771.
The inflow angle of the measurements has been calculated from the mean value of
the wind direction in each (6.5 h) time series, and used to determine the decay factor
for the coherence according to Eq. (19) for each of the six simulations. The
corresponding coherence (‘‘Davenport’’) is also indicated in the figure as a straight
line.
The coherence comparisons show very good agreement between measurements
and simulations, indicating that the Davenport type coherence and Schlez and
Infield’s decay factors are reasonable in this case.
Finally, the simulated and measured cross correlation functions are compared in
Fig. 8. 8  10 min with average inflow angles a ¼ 70:51 have been found in the
6  6.5 h wind speed time series and used to estimate the cross correlation functions.
This is not enough data to estimate the coherence in this narrow wind direction
P. Srensen et al. / J. Wind Eng. Ind. Aerodyn. 90 (2002) 1381–1402 1397

100
Measured
Simulated
sqrt(coherence) Davenport

10-1
0 0.001 0.002 0.003 0.004 0.005 0.006 0.007 0.008 0.009 0.01
frequency [Hz]
Fig. 7. Square root coherence function of measured and simulated 6  6.5 h wind speed time series on
Vindeby SMW and SMS.

1
Measured
0.9 Simulated
Normalised cross correlation function

Sim-frozen
0.8

0.7

0.6

0.5

0.4

0.3

0.2

0.1

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Normalised delay
Fig. 8. Measured and simulated cross correlation functions of the eight 10 min time series with average
inflow angles a ¼ 70:51: Simulation is done with the present model, whereas Sim frozen is done assuming
Taylor’s frozen turbulence.

band, but as it can be seen from Fig. 8, the estimated cross correlation functions are
reasonably smooth.
The exactly measured inflow angle has been used in the simulations, ensuring that
also distributed inflow angles in the interval a ¼ 70:51 for the simulations. Two
1398 P. Srensen et al. / J. Wind Eng. Ind. Aerodyn. 90 (2002) 1381–1402

simulations have been done for each of the time series, one with the present
PARKWIND model using the Schlez and Infield decay factors along and alat ; and one
assuming Taylor’s frozen turbulence. The frozen turbulence is also simulated with
PARKWIND, using the longitudinal decay factor along ¼ 0:
The values shown in Fig. 8 have been normalised, so that the delay=1
corresponds to the travel time of the wind form SMW to SMS, using the mean
wind speed in each of the 10 min time series. The cross correlation functions are
normalised with the standard deviations of wind speeds, so that maximum=1
corresponds to identical, only time delayed time series at the two masts. The shown
curves are the averages of the eight cross correlation functions.
The comparison of the measured and simulated cross correlation functions in
Fig. 8 shows a good agreement between our simulations and the measurements, and
also reveals the problem with the assumption of frozen turbulence. It has a very high
spike, because the turbulence at the two masts are assumed to be (almost) identical,
only with a delay corresponding to the travel time from SMW to SMS.
The rotor wind model is difficult to verify, because the equivalent wind speed is a
fictitious concept, which cannot be measured directly. Still, Rosas [18] compared
simulations of the equivalent wind speed to measurements of the wind speed
performed by a pitot tube mounted on a rotating wind turbine blade. These
measurements are described in Petersen and Madsen [23].
The undisturbed wind speed measured at hub height on a meteorological mast in
front of the wind turbine was used to determine the mean wind speed and turbulence
intensity as input parameters to the simulations. The turbulence intensity was 0.16,
whereas the mean wind speed was 10.5 m/s.
The pitot tube was mounted in 15 m radius. The main parameters of the wind
turbine are a rotor diameter of 41 m, a rotor speed of 30 rpm, a distance from the
rotor disc to the tower of 2.9 m and a tower diameter of 1.7 m. These values were also
used as input to the simulations.
The measured wind speed is seen from a single point on a rotating blade, whereas
the rotor wind model describes the summed effect of the wind speed on all
three blades. Therefore, the measured wind speed was processed to reflect the
average wind speed from 3 blades before it was compared to the simulation. This was
done by averaging the measured wind speed at times tDt, t and t+Dt, where Dt is
the time corresponding to 1201 rotation with 30 rpm.
The comparison of the simulation with the processed measurements is shown in
Fig. 9. The comparison between the measured and simulated wind speeds shows
significantly higher measurement values than simulation values around the 3p
frequency (1.5 Hz). The main reason for that difference is that the measurements are
done in a single radius, whereas the model intends to include the averaging along the
blades.
One exception from that is the spike on the simulated wind speed exactly at the 3p
frequency, which is probably due to overestimation of the tower shadow effects.
Another significant difference between measurements and simulations in Fig. 9 is
the 6p on the measurements, which we do not simulate. The 6p could easily be
included in the simulations, but the reason why we have not included is that it has
P. Srensen et al. / J. Wind Eng. Ind. Aerodyn. 90 (2002) 1381–1402 1399

0.20
Measured
0.18 Simulated
0.16
0.14
PSD [(m/s)2/Hz]

0.12
0.10
0.08
0.06
0.04
0.02
0.00
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
Frequency [Hz]
Fig. 9. PSDs of simulated equivalent wind speed and measured wind speed measured with a pitot tube on
a rotating blade.

12 March 2001 18:17


Measured
Measured
Simulated
Simulated
106
Power PSD [(kW) /Hz]
2

104

102

100

10-2 10-1 100 101


Frequency [Hz]
Fig. 10. PSDs of measured and simulated power of a 2 MW NM2000/72 wind turbine.

only little influence on the power output from the wind turbines, because the wind
turbine structures act as a low-pass filter.
This statement can be confirmed by comparisons of measured and simulated
power from a wind turbine. Fig. 10 shows the PSDs of measured and simulated
power on a 2 MW NM2000/72 NEG-Micon wind turbine.
1400 P. Srensen et al. / J. Wind Eng. Ind. Aerodyn. 90 (2002) 1381–1402

Here we can also see the concentration of energy around the 3p frequency, which
is 0.9 Hz in this case. The 6p; 9p and 12p frequencies can also be seen in the
measurements, but they are much weaker than the 3p:
In Fig. 10, the reduction of the measured 6p compared to the 3p is much more
significant. Of course, it is different wind turbines and Fig. 10, but the difference also
reflects the general aspect that the mechanical structure of the wind turbine low-pass
filters the wind speed. Consequently, the higher p’s are generally more significant in
the wind speed than in the power.

6. Conclusion

Models for the wind speeds in wind farms have been developed. The models
include the effects which are assumed to be important to predict the influence of the
wind farm on the power quality as characterised in a new IEC standard for power
quality of wind turbines.
The selected effects can be summarised as the park scale coherence between the
wind speeds at different wind turbines in the farm, and the effects of rotational
sampling in the rotor. Another way to classify the effects is as turbulence effects and
tower shadow effects.
Other effects like wind shear and yaw errors, which are important for the
structural loading of wind turbine blades, have not been included here. This is
because the summation of the torque from the blades removes most of these effects.
The models have been verified using a few sets of data, and the results are very
promising. The park scale model is able to simulate the delay of the coherent part of
the wind speed between two points with longitudinal separation, and the coherence
between the simulated wind speeds shows good agreement with the corresponding
coherence between measured wind speeds. Comparisons of the cross correlation
functions have shown that the present model is more reliable than models assuming
frozen turbulence, because we include the decay in coherence for longitudinal
separation.
The rotor wind model is more difficult to verify directly, because the eq-
uivalent wind speed generated by this model is a fictitious concept, which cannot
be measured directly. However, the equivalent wind speed has been compared
to wind speeds measured with a pitot tube mounted on a rotating wind turbine
blade. The result of this comparison is reasonable up to approximately four
times the rotor speed, taking into account the difference between the (measured)
single radius wind speed and the (simulated) radius weighted equivalent wind
speed.
The rotor wind model has also been verified indirectly by a comparison of the
PSDs of measured and simulated power. This comparison also shows good
agreement for frequencies up to four times the rotor speed, and it demonstrates that
the structure of the wind turbine acts as a filter, which reduces the fluctuations at
higher frequencies.
P. Srensen et al. / J. Wind Eng. Ind. Aerodyn. 90 (2002) 1381–1402 1401

Acknowledgements

The Danish Energy Agency is acknowledged for the funding of this work in
contract 1363/00-0003. Besides, a special thanks is given to our partners in the
Danish project, Aalborg University and Dancontrol Engineering and North-West
Sealand Energy Supply Company, NVE. Also thanks to the Danish transmission
system operators Eltra and Elkraft System, who have participated as members of an
advisory committee for the project. Finally thanks to SEAS Wind Energy Centre for
funding of maintenance of Vindeby measurements. Finally, Capes is acknowledged
for funding the Brazilian Ph.D. project.

References

[1] P. Srensen, A.D. Hansen, L. Janosi, B. Bak-Jensen, F. Blaabjerg, J. Bech, Simulation of wind farm
interaction with grid, 2001 European Wind Energy Conference and Exhibition, 2–6 July 2001,
Copenhagen, DK.
[2] IEC 61400-21, Ed.1: Wind turbine generator systems—Part 21: measurement and assessment of
power quality characteristics of grid connected wind turbines, Final Draft International Standard
88/144/FDIS International Electrotechnical Commission, IEC 2001-07-01.
[3] L. Kristensen, S. Frandsen, Model for power spectra of the blade of a wind turbine measured from
the moving frame of reference, J. Wind Eng. Ind. Aerodyn. 10 (1982).
[4] J. Tande, P. Srensen, M. Galal, U. Naoman, E. Mansour, A. Hewehy, A. Maghawry, I. Darweesh,
Hurghada Wind Energy Technology Centre Demostration Wind Farm Studies, Power Quality
Assessment, Ris-I-1141(EN), September 1997.
[5] P. Srensen, Methods for calculation of the flicker contribution from wind turbines Ris-I-939, 1995,
Ris National Laboratory, Roskilde, Denmark.
[6] N.O. Jensen, Anote on wind generator interaction, Ris M 2411, November 1983.
[7] S. Frandsen, M.L. Thgersen, Integrated fatigue loading for wind turbines in wind farms by
combining ambient turbulence and wakes, Wind Eng. 23 (6) (1999) 327–340.
[8] A.I. Estanqueiro, R.F. Aguiar, J.A.G. Saraiva, R.M.G. Castro, J.M.F.D. Jesus, On the effect of
utility grid characteristics on wind park power output fluctuations, Proceedings of British Wind
Energy Conference BWEA’15, York, 1993.
[9] M. Shinozuka, C.M. Jan, Digital simulations of random processes and its applications, J. Sound Vib.
25 (1) (1972) 111–128.
[10] J. Mann, Wind field simulation, Prob. Eng. Mech. 13 (4) (1998) 269–282.
[11] P. Srensen, B. Bak-Jensen, J. Kristiansen, A.D. Hansen, L. Janosi, J. Bech, Power plant
characteristics of wind farms, in: Proceedings (on DC-ROM), Wind Power for the 21 Century:
EUWEC Special Topic Conference and Exhibition, 25–27 September 2000, Kassel, DE (WIP—
.
Renewable Energies, Munchen, 2001) pp. 176–179.
[12] J.C. Kaimal, J.C. Wyngaard, Y. Izumi, O.R. Cot!e, Spectral characteristics of surface layer
turbulence, Q. J. R. Meteorol. Soc. 98 (1972) 563–598.
[13] J. Hjstrup, S.E. Larsen, P.H. Madsen, Power spectra of horizontal wind components in the neutral
atmospheric boundary layer, in: N.O. Jensen, L. Kristiansen, S.E. Larsen (Eds.), Ninth Symposium
of Turbulence and Diffusion, American Meteorology Society, 1990, pp. 305–308.
[14] Dansk Ingenirforenings og Ingenir-sammenslutningens norm for last og sikkerhed for
vindmllekonstruktioner. 1. udgave maj 1992. Dansk Standard DS 472.
[15] A.G. Davenport, The spectrum of horizontal gustiness near the ground in high winds, Q. J. R.
Meteorol. Soc. 87 (1961) 194–211.
[16] W. Schlez, D. Infield, Horizontal, two point coherence for separations greater than the measurement
height, Boundary-Layer Meteor. 87 (1998) 459–480.
1402 P. Srensen et al. / J. Wind Eng. Ind. Aerodyn. 90 (2002) 1381–1402

[17] W. Langreder, Models for Variable Speed Wind Turbines, M.Sc. Thesis, CREST Loughborough
University and Ris National Laboratory, 1996.
[18] P.A.C. Rosas, P. Srensen, H. Bindner, Fast wind modelling for wind turbines, in: Proceedings
(on CD-ROM), Wind power for the 21 Century: EUWEC Special Topic Conference and Exhibition,
.
25–27 September 2000, Kassel, DE, (WIP—Renewable Energies, Munchen, 2001) pp. 184–187.
[19] P.H. Madsen, F. Rasmussen, Rotor loading on a three-bladed wind turbine, European Wind Energy
Conference EWEC’89, Glasgow, 1989.
[20] P. Srensen, G.C. Larsen, C.J. Christensen, A complex frequency domain model of wind turbine
structures, J. Sol. Energy Eng. 117 (1995) 311–317.
[21] P. Srensen, Frequency domain modelling of wind turbine structures. Ris-R749(EN), 1994.
[22] F.A. Ekelund, Hydrodynamik, Newtonske v!cskers mekanik. (Den private Ingenirfond). 322pp.
1968.
[23] J.T. Petersen, H.Aa. Madsen, Local inflow and dynamics—measured and simulated on a rotating
wind turbine blade, DRAFT, Ris-R-993(EN), September 1997.

You might also like