Sie sind auf Seite 1von 46

ARTICLE IN PRESS

Progress in Aerospace Sciences 42 (2006) 285330


State of the art in wind turbine aerodynamics and aeroelasticity
M.O.L. Hansen
a,
, J.N. Srensen
a
, S. Voutsinas
b
, N. Srensen
c,d
, H.Aa. Madsen
c
a
Department of Mechanical Engineering, Technical University of Denmark, Fluids Section, Nils Koppels Alle,
Building 403, DK-2800 Lyngby, Denmark
b
Department of Mechanical Engineering, National Technical University of Athens, Fluids Section, 15780 Zografou, Greece
c
Wind Energy Department, Risoe National Laboratory, Building VEA-762, P.O. Box 49, Frederiksborgvej 399, DK-4000 Roskilde, Denmark
d
Department of Civil Engineering, Alborg University, Sohngaardsholmsvej 57, DK 9000 Aalborg, Denmark
Available online 29 December 2006
Abstract
A comprehensive review of wind turbine aeroelasticity is given. The aerodynamic part starts with the simple
aerodynamic Blade Element Momentum Method and ends with giving a review of the work done applying CFD on wind
turbine rotors. In between is explained some methods of intermediate complexity such as vortex and panel methods. Also
the different approaches to structural modelling of wind turbines are addressed. Finally, the coupling between the
aerodynamic and structural modelling is shown in terms of possible instabilities and some examples.
r 2006 Elsevier Ltd. All rights reserved.
Keywords: Aeroelasticity; Wind turbines
Contents
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 286
2. Predicting aerodynamic loads on a wind turbine. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287
2.1. Blade Element Momentum Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287
2.1.1. Dynamic wake/inow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 288
2.1.2. Yaw/tilt model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 289
2.1.3. Dynamic stall . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 289
2.1.4. Airfoil data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 290
2.1.5. Wind simulation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 290
2.2. Lifting line, panel and vortex models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 291
2.2.1. Vortex methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 291
2.2.2. Panel methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 292
2.3. Generalized actuator disc models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295
2.4. NavierStokes solvers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 297
2.4.1. Introduction to computational rotor aerodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 297
2.4.2. Approaches . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 298
2.4.3. Turbulence and transition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 299
www.elsevier.com/locate/paerosci
0376-0421/$ - see front matter r 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.paerosci.2006.10.002

Corresponding author. Tel.: +45 45254316.


E-mail address: molh@mek.dtu.dk (M.O.L. Hansen).
2.4.4. Geometry and grid generation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 299
2.4.5. Numerical issues. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 299
2.4.6. Application of CFD to wind turbine aerodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 300
2.4.7. Future . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 302
3. Structural modelling of a wind turbine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 302
3.1. Principle of virtual work and use of modal shape functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 302
3.2. FEM modelling of wind turbine components applying non-linear beam theory. . . . . . . . . . . . . . . . . . 304
4. Problems and solutions in wind turbine aeroelasticity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 308
4.1. Aeroelastic stability. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 308
4.2. Aeroelastic coupling: linear vs. non-linear formulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 309
4.3. Examples of time simulations and instabilities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 310
4.3.1. Edgewise blade vibration instability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 312
4.3.2. Instability problems of parked rotors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 317
4.3.3. Flutter instability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 317
5. Present and future developments of aeroelastic models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 318
5.1. Areas with inuence on the development of aeroelastic models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 318
5.1.1. Inuence of up-scaling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 318
5.1.2. Siting of the turbines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319
5.1.3. Future trends in turbine design and siting. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319
5.2. Areas of development in present and new codes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319
5.2.1. Non-linear structural dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319
5.2.2. Calculation of induction and its dynamics. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 320
5.2.3. Wake operation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 321
5.2.4. Derivation of airfoil data for aeroelastic simulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 322
5.2.5. Complex inow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 323
5.2.6. Aerodynamics of parked rotors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 324
5.2.7. Offshore turbines including oating turbines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 324
6. Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 325
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 325
1. Introduction
The size of commercial wind turbines has
increased dramatically in the last 25 years from
approximately a rated power of 50 kW and a rotor
diameter of 1015 m up to todays commercially
available 5 MW machines with a rotor diameter of
more than 120 m. This development has forced the
design tools to change from simple static calcula-
tions assuming a constant wind to dynamic simula-
tion software that from the unsteady aerodynamic
loads models the aeroelastic response of the entire
wind turbine construction, including tower, drive
train, rotor and control system. The Danish
standard DS 472 [1] allows simplied load calcula-
tions if the rotor diameter is less than 25 m and
some other criteria are fullled. A rotor diameter of
25 m corresponds approximately to a rated power of
200250 kW, which is less than almost any modern
commercial wind turbine today. Instead, modern
wind turbines are designed to fulll the require-
ments of the more comprehensive IEC 61 400-1 [2]
standard. At some time during the development of
larger and larger commercial wind turbines the need
for aeroelastic tools thus became necessary. Aero-
elastic tools were mainly developed at the univer-
sities and research laboratories in parallel with the
evolution of commercial wind turbines. At the same
time governments and utility companies erected
large non-commercial prototypes for research pur-
poses, as the Nibe [3] and Tjaereborg machines [4].
Measurement campaigns were undertaken on these
machines and the results used to tune and validate
the aeroelastic programmes, in order to develop
advanced software for the rapidly growing industry.
Even today measurements from the Tjaereborg
machine is used as a benchmark when developing
new aeroelastic codes, see e.g. [5]. In [5] is also
compiled a long list of available software that at
different levels of complexity can model the aero-
elastic response of a wind turbine construction. All
the aeroelastic models need as input a time history
of the wind seen by the rotor, which as a minimum
must contain some physical correct properties such
as realistic power spectra and spatial coherence.
Apart from the wind input aeroelastic codes contain
ARTICLE IN PRESS
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 286
an aerodynamic part to determine the wind loads
and a structural part to describe the dynamic
response of the wind turbine construction. For the
aerodynamic part most codes use the Blade Element
Momentum Method (BEM) as described by
Glauert [6], since this method is very fast and,
provided that reliable airfoil data exist, yields
accurate results. Therefore, this method, with all
the necessary engineering adds on, is thoroughly
described later in this article. However, more
advanced numerical models based on the Euler
and NavierStokes (NS) equations are becoming so
fast that they now begun to replace the BEM
method in some situations, e.g. when analysing yaw
or interaction between wind turbines in parks.
These models contain more physics and less
empirical input than the BEM method and are
extensively described in this paper. The discretiza-
tion of the wind turbine structure is presently where
the various available codes differ most. Roughly,
there exist three ways to model the structural
dynamics of a wind turbine. One is a full Finite
Element Method (FEM) discretization and another
is a multi-body formulation, where different rigid
parts are connected through springs and hinges.
Finally, the description of blade and tower deec-
tions can be made as a linear combination of some
physical realistic modes; typically the lowest eigen-
modes. The last method greatly reduces the
computational time per time step, as compared
with a full FEM discretization. All the various ways
of discretizing the wind turbine structure will be
treated in details later in the paper. The very
detailed description of the aerodynamic and struc-
tural models is where this paper differs mostly from
other review articles concerning wind turbine
aeroelasticity such as e.g. [79].
2. Predicting aerodynamic loads on a wind turbine
Methods of various levels of complexity to
calculate the aerodynamic loads on a wind turbine
rotor are given, starting with the popular BEM, and
ending with the solution of the NS equations.
2.1. Blade Element Momentum Method
BEM is the most common tool for calculating the
aerodynamic loads on wind turbine rotors since it is
computationally cheap and thus very fast. Further,
it provides very satisfactory results provided that
good airfoil data are available for the lift and drag
coefcients as a function of the angle of attack, and
if possible, the Reynolds number. The method was
introduced by Glauert [6] as a combination of one-
dimensional (1D) momentum theory and blade
element considerations to determine the loads
locally along the blade span. The method assumes
that all sections along the rotor are independent and
can be treated separately; typically in the order of
1020 radial sections are calculated. At a given
radial section a difference in the wind speed is
generated from far upstream to deep in the wake.
The resulting momentum loss is due to the axial
loads produced locally by the ow passing the
blades, creating a pressure drop over the blade
section. The local angle of attack at a given radial
section on a blade can be constructed, provided that
the induced velocity generated by the action of the
loads is known, see Fig. 1. V
0
is the undisturbed
wind velocity, W, the induced velocity, V
rot
= o r
the rotational speed of the blade section, V
blade
the
velocity of the blade section apart from the blade
rotation and b is the local angle of the blade section
to the rotor plane.
Combining the global momentum loss with the
loads generated locally at the blade section yields
formulas for the induced velocity as
W
z
=
BLcos f
4rprF V
0
f
g
n(n W)

, (2.1.1)
W
y
=
BLsin f
4rprF V
0
f
g
n(n W)

, (2.1.2)
B is the number of blades, L the lift computed from
the lift coefcient, f is the ow angle, r the density
of air, r the radial position considered, V
0
the wind
velocity, W the induced velocity and n the normal
vector to the rotor plane. F is Prandtls tip loss
correction that corrects the equations to be valid for
a nite number of blades, see [6, 10]. If there is no
yaw misalignment, that is, the normal vector to the
rotor plane, n, is parallel to the wind vector, then
ARTICLE IN PRESS
W
V
o
V
blade
V
rot
V
rel
y x
rotor plane
z

Fig. 1. Construction of angle of attack, a.


M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 287
Eq. (2.1.1) reduces to the well-known expression
C
T
= 4aF(1 f
g
a), (2.1.3)
where by denition for an annual element of
innitesimal thickness, dr, and area, dA = 2pr dr,
C
T
=
dT
1=2rV
2
0
dA
. (2.1.4)
The axial interference factor is dened as
a =
W
z
V
0
(2.1.5)
and f
g,
usually referred to as the Glauert correction,
is an empirical relationship between C
T
and a, in the
turbulent wake state. It may assume the form
f
g
=
1 for ap0:3;
1
4
(5 3a) for a40:3:
_
(2.1.6)
Eqs. (2.1.1) and (2.1.2) are also known to be valid
for an extreme yaw misalignment of 901, that is, the
incoming wind is parallel to the rotor plane as a
helicopter in forward ight. Without any proof
Glauert therefore assumed that Eqs. (2.1.1) and
(2.1.2) are valid for any yaw angle.
An aeroelastic code is running in the time domain
and for every time step the aerodynamic loads must
be calculated at all the chosen radial stations along
the blades as input to the structural model. For a
given time the local angle of attack is determined on
every point on the blades, as indicated in Fig. 1. The
lift and drag coefcients can now be found from
table look-up, and the lift can be determined.
The induced velocities can now be updated using
Eqs. (2.1.1) and (2.1.2) simply assuming old values
for the induced velocities on the right-hand sides
(RHS). Updating the RHS of Eqs. (2.1.1) and
(2.1.2) could continue until the equations are solved
with all values at the same time step. However, this
is not necessary as this update takes place in the
next time step, i.e. time acts as iteration. More
important, the values of the induced velocities
change very slowly in time due to the phenomena
of dynamic inow or dynamic wake.
2.1.1. Dynamic wake/inow
The induced velocities calculated using Eqs. (2.1.1)
and (2.1.2) are quasi-steady, in the sense that they
give the correct values only when the wake is in
equilibrium with the aerodynamic loads. If the loads
are changed in time there is a time delay proportional
to the rotor diameter divided by the wind speed
before a new equilibrium is achieved. To take into
account this time delay, a dynamic inow model
must be applied. In two EU-sponsored projects
([11,12]) different engineering models were tested
against measurements. One of these models, pro-
posed by S. ye, is a lter for the induced velocities
consisting of two rst-order differential equations
W
int
t
1
dW
int
dt
= W
qs
k t
1
dW
qs
dt
, (2.1.7)
W t
2
dW
dt
= W
int
, (2.1.8)
W
qs
is the quasi-static value found by Eqs. (2.1.1)
and (2.1.2), W
int
an intermediate value and W the
nal ltered value to be used as the induced velocity.
The two time constants are calibrated using a simple
vortex method as
t
1
=
1:1
(1 1:3a)

R
V
0
(2.1.9)
and
t
2
= 0:39 0:26
r
R
_ _
2
_ _
t
1
, (2.1.10)
where R is rotor radius.
In Fig. 2 is shown for the Tjaereborg machine the
computed and measured response on the rotorshaft
torque for a sudden change of the pitch angle. At
t = 2 s the pitch is increased from 01 to 3.71,
decreasing the local angles of attack. First the
rotorshaft torque drops from 260 to 150 kNm and
not until approximately 10 s later the induced
velocities and thus the rotorshaft torque have settled
ARTICLE IN PRESS
400
350
300
250
200
150
100
0 10 20 30 40 50 60
time [s]
R
o
t
o
r
s
h
a
f
t

t
o
r
q
u
e

[
k
N
m
]
BEM
Measurement
Fig. 2. Comparison between measured and computed time series
of the rotorshaft torque for the Tjaereborg machine during a step
input of the pitch for a wind speed of 8.7 m/s.
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 288
at a new equilibrium. At t = 32 s the pitch is
changed back to 01 and a similar overshoot in
rotorshaft torque is observed. The decay of the
spikes seen in Fig. 2 can only be computed with a
dynamic inow model, and such a model is there-
fore of utmost importance for a pitch-regulated
wind turbine.
2.1.2. Yaw/tilt model
Another engineering model for the induced
velocities concerns yaw or tilt. When the rotor disc
is not perfectly aligned with the incoming wind there
is an angle different from zero between the rotor
normal vector and the incoming wind, see Fig. 3. A
yaw/tilt model redistributes the induced velocities so
that the induced velocities are higher when a blade is
positioned deep in the wake than when it is pointing
more upstream. An example of such a model, taken
from helicopter literature [13] is given below. Here,
the input is the induced velocity, W
0
, calculated
using Eqs. (2.1.1), (2.1.2), (2.1.7) and (2.1.8). The
output is a redistributed value, nally used when
estimating the local angle of attack, W,
W = W
0
1
r
R
tan
w
2
cos(y
b
y
0
)
_ _
, (2.1.11)
y
b
is the actual position of a blade, y
0
, is the position
where the blade is furthest downstream and w is the
wake skew angle, see Fig. 3. In some BEM
implementations, W
0
, is the average value of all
blades at the same radial position, r, and in other
codes it is the local value. This difference in
implementation may cause a small difference from
code to code. Further, there exist different mod-
ications of Eq. (2.1.11) from different codes, see
[12]. A yaw/tilt model increases the induced
velocities on the downstream part of the rotor and
decreases similarly the induced velocity on the
upstream part of the rotor disc. This introduces a
yaw moment that tries to align the rotor with the
incoming wind, hence tending to reduce yaw
misalignment. For a free yawing turbine such a
model is, therefore, of utmost importance when
estimating the yaw stability of the machine.
2.1.3. Dynamic stall
The wind seen locally on a point on the blade
changes constantly due to wind shear, yaw/tilt
misalignment, tower passage and atmospheric
turbulence. This has a direct impact on the angle
of attack that changes dynamically during the
revolution. The effect of changing the blades angle
of attack will not appear instantaneously but will
take place with a time delay proportional to the
chord divided with the relative velocity seen at the
blade section. The response on the aerodynamic
load depends on whether the boundary layer is
attached or partly separated. In the case of attached
ow the time delay can be estimated using
Theodorsen theory for unsteady lift and aerody-
namic moment [14]. For trailing edge stall, i.e. when
separation starts at the trailing edge and gradually
increases upstream at increasing angles of attack,
so-called dynamic stall can be modelled through a
separation function, f
s
, as described in [15], see later.
The BeddoesLeishman model [16] further takes
into account attached ow, leading edge separation
and compressibility effects, and also corrects the
drag and moment coefcients. For wind turbines,
trailing edge separation is assumed to represent the
most important phenomenon regarding dynamic
airfoil data, but also effects in the linear region may
be important, see [17]. It is shown in [15] that if a
dynamic stall model is not used one might compute
apwise vibrations, especially for stall regulated
wind turbines, which are non-existing on the real
machine. For stability reasons it is thus highly
recommended to at least include a dynamic stall
model for the lift. For trailing edge stall the degree
of stall is described through f
s
, as
C
l
(a) = f
s
C
l;inv
(a) (1 f
s
)C
l;fs
(a), (2.1.12)
where C
l,inv
denotes the lift coefcient for inviscid
ow without any separation and C
l,fs
is the lift
coefcient for fully separated ow, e.g. on a at
plate with a sharp leading edge. C
l,inv
is normally an
extrapolation of the static airfoil data in the linear
ARTICLE IN PRESS
Rotor disc
V
o
V
o
n
x

yaw/tilt
V
,
W
n
Fig. 3. Wind turbine rotor not aligned with the incoming wind.
The angle between the velocity in the wake (the sum of the
incoming wind and the induced velocity normal to the rotor
plane) is denoted the wake skew angle, w.
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 289
region, and in [17] a way of estimating C
l,fs
and f
s
st
is
shown. f
s
st
is the value of f
s
that reproduces the static
airfoil data when applied in Eq. (2.1.12). The
assumption is that, f
s
, always will try to get back
to the static value as
df
s
dt
=
f
st
s
f
s
t
, (2.1.13)
that can be integrated analytically to give
f
s
(t Dt) = f
st
s
(f
s
(t) f
st
s
) exp(Dt=t), (2.1.14)
t is a time constant approximately equal to Ac/V
rel
,
where c denotes the local chord, and V
rel
is the
relative velocity seen by the blade section. A is a
constant that typically takes a value about 4.
Applying a dynamic stall model the airfoil data is
always chasing the static value at a given angle of
attack that is also changing in time. If e.g. the angle
of attack is suddenly increased from below to above
stall the unsteady airfoil data contains for a short
time some of the inviscid/unstalled value, C
l,inv
, and
an overshoot relative to the static data is seen. It can
thus been seen as a model of the time constant for
the viscous boundary layer to develop from one
state to another.
2.1.4. Airfoil data
The BEM as described above, including all
engineering corrections, is used in most aeroelastic
codes to compute the unsteady aerodynamic loads
on wind turbine rotors. The method is often quite
successful, but depends on reliable airfoil data for
the different blade sections. Three-dimensional (3D)
effects from the tip vortices are taken into account
when applying Prandtls tip loss correction and after
this correction the local ow around a given blade
section is assumed to be two-dimensional, i.e. 2D
airfoil data from wind tunnel measurements are
used. However, such measurements are often
limited to the maximum lift coefcient, C
l,max
for
airplanes that usually are operated at unstalled ow
conditions. Further, at higher values it is difcult to
measure the forces because of the unsteady and 3D
nature of stall. In contrast to airplane wings, a wind
turbine blade often operates in deep stall, especially
for stall regulation. For the inner part of the blades
even data for low angles of attack might be difcult
to nd in literature since for structural reasons the
airfoils used are much thicker than those used on
airplanes. Further, because of rotation the bound-
ary layer is subjected to Coriolis- and centrifugal
forces, which alter the 2D airfoil characteristics.
This is especially pronounced in stall. It is thus often
necessary to extrapolate existing airfoil data into
deep stall and to include the effect of rotation.
Methods have been developed that from a CFD
calculation of the ow past a full wind turbine rotor
can extract 3D airfoil data [18], which then later can
be applied in aeroelastic calculations using the much
faster BEM method. In this method, the induced
velocity at the rotor plane is estimated from the
azimuthally averaged velocity in very thin annular
elements up- and downstream of the rotorplane. In
[19,20], two engineering methods to correct 2D
airfoil data for 3D rotational effects are given as
C
x;3D
= C
x;2D
a(c=r)
h
cos
n
b DC
x
; x = l; d; m;
(2.1.15)
DC
l
= C
l;inv
C
l;2D
,
DC
d
= C
d;2D
C
d;2Dmin
DC
m
= C
m;2D
C
m;inv
,
c is the chord, r the radial distance to rotational axis
and b the twist.
In [19] only the lift is corrected, i.e. x = l, and the
constants are a = 3, n = 0 and h = 2, whereas in [20]
a = 2.2, n = 4 and h = 1. In [21] another method
based on correcting the pressure distribution along
the airfoil is given. One must, however, be very
aware that the choice of airfoil data directly
inuences the results from the BEM method. For
certain airfoils a lot of experience has been gathered
regarding appropriate corrections to be used in
order to obtain good results, and, because of this,
blade designers tend to be conservative in their
choice of airfoils. With maturing CFD algorithms
especially for the transition and turbulence models
and more wind tunnel tests, the trend is now to use
airfoils specially designed and dedicated to wind
turbine blades, see e.g. [22].
2.1.5. Wind simulation
Besides airfoil data, also realistic spatial-temporal
varying wind elds must be generated as input to an
aeroelastic calculation of a wind turbine. As a
minimum the simulated eld must satisfy some
statistical requirements, such as a specied power
spectre and spatial coherence, see [23,24]. In this
method, each velocity component is generated
independently from the others, meaning that there
is no guarantee for obtaining correct cross-correla-
tions. In [25], a method ensuring this is developed
ARTICLE IN PRESS
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 290
on the basis of the linearized NS equations. In the
future, wind elds are expected to be generated
numerically from Large Eddy Simulations (LES) or
Direct Numerical Simulations (DNS) of the NS
equations for the ow on a landscape similar to the
actual siting of a specic wind turbine.
2.2. Lifting line, panel and vortex models
In the present section, 3D inviscid aerodynamic
models are reviewed. They have been developed in
an attempt to obtain a more detailed description of
the 3D ow that develops around a wind turbine.
The fact that viscous effects are neglected is
certainly restrictive as regards the usage of such
models on wind turbines. However, they should be
given the credit of contributing to a better under-
standing of dynamic inow effects as well as the
credit of providing a better insight into the overall
ow development [11,12]. There have been attempts
to introduce viscous effects using viscousinviscid
interaction techniques [26,27] but they have not yet
reached the required maturity so as to become
engineering tools, although they are full 3D models
that can be used in aeroelastic analyses.
2.2.1. Vortex methods
In vortex models the rotor blades, trailing and
shed vorticity in the wake are represented by lifting
lines or surfaces [28]. On the blades the vortex
strength is determined from the bound circulation
that stems from the amount of lift created locally by
the ow past the blades. The trailing wake is
generated by the spanwise variation of the bound
circulation while the shed wake is generated by a
temporal variation and ensures that the total
circulation over each section along the blade
remains constant in time. Knowing the strength
and position of the vortices the induced velocity
can be found in any point using the BiotSavart
law, see later. In some models (namely the lifting-
line models) the bound circulation is found from
airfoil data table-look up just as in the BEM
method. The inow is determined as the sum of
the induced velocity, the blade velocity and
the undisturbed wind velocity, see Fig. 1. The
relationship between the bound circulation and the
lift is denoted as the KuttaJoukowski theorem
(rst part of Eq. (2.2.1)), and using this together
with the denition of the lift coefcient (second
part of Eq. (2.2.1)) a simple relationship between
the bound circulation and the lift coefcient can
be derived
L = rV
rel
G = 1=2rV
2
rel
cC
l
= G = 1=2V
rel
cC
l
.
(2.2.1)
Any velocity eld can be decomposed in a
solenoidal part and a rotational part as
V = V C VF, (2.2.2)
where C is a vector potential and F a scalar
potential [29]. From Eq. (2.2.2) and the denition of
vorticity a Poisson equation for the vector potential
is derived
V
2
C = o. (2.2.3)
In the absence of boundaries, C, can be expressed in
convolution form as
C(x) =
1
4p
_
o
/
x x
/
[ [
dVol, (2.2.4)
where x denotes the point where the potential is
computed, a prime denotes evaluation at the point
of integration x
/
which is taken over the region
where the vorticity is non-zero, designated by Vol.
From its denition the resulting induced velocity
eld is deduced from the induction law of Biot
Savart
w(x) =
1
4p
_
(x x
/
) o
/
x x
/
[ [
3
dVol. (2.2.5)
In its simplest form the wake from one blade is
prescribed as a hub vortex plus a spiralling tip
vortex or as a series of ring vortices. In this case the
vortex system is assumed to consist of a number of
line vortices with vorticity distribution
o(x) = Gd(x x
/
), (2.2.6)
where G is the circulation, d is the line Dirac delta
function and x
/
is the curve dening the location of
the vortex lines. Combining (2.2.5) and (2.2.6)
results in the following line integral for the induced
velocity eld:
w(x) =
1
4p
_
S
G
(x x
/
)
x x
/
[ [
3

qx
/
qS
/
qS
/
, (2.2.7)
where S is the curve dening the vortex line and S
/
is
the parametric variable along the curve.
Utilizing (2.2.7) simple vortex models can be
derived to compute quite general ow elds about
wind turbine rotors. The rst example of a simple
vortex model is probably the one due to Joukowski
[30], who proposed to represent the tip vortices by
an array of semi-innite helical vortices with
ARTICLE IN PRESS
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 291
constant pitch, see also [31]. In [32], a system of
vortex rings was used to compute the ow past a
heavily loaded wind turbine. It is remarkable that in
spite of the simplicity of the model, it was possible
to simulate the vortex ring/turbulent wake state
with good accuracy, as compared to the empirical
correction suggested by Glauert, see [6]. Further, a
similar simple vortex model was used in [33] to
calculate the relation between thrust and induced
velocity at the rotor disc of a wind turbine, in order
to validate basic features of the streamtube-
momentum theory. The model includes effects of
wake expansion, and, as in [32] simulates a rotor
with an innite number of blades, with the wake
being described by vortex rings. From the model it
was found that the axial induced velocities at the
rotor disc are smaller than those determined from
the ordinary stream tube-momentum theory. A
similar approach has been utilized by Koh and
Wood [34] and Wood [35] for studying rotors
operating at high tip speed ratios.
2.2.2. Panel methods
The inviscid and incompressible ow past the
blades themselves can be found by applying a
surface distribution of sources s and dipoles m
(Fig. 4). The background is Greens theorem, which
allows obtaining an integral representation of any
potential ow eld in terms of the singularity
distribution [36,37].
V(x; t) = V
0
Vf(x; t):
f(x; t) =
_
S
s
/
(t)
4p x x
/
[ [

m
/
(t) n
/
:r
4p x x
/
[ [
3
_ _
dS
/
,
(2:2:8)
V
0
denotes a given (external) potential ow eld
possibly varying in time and space and f is the
perturbation scalar potential. S stands for the active
boundary of the ow and includes the solid
boundaries of the ow S
B
as well as the wake
surfaces of all lifting components S
W
. In (2.2.8) m, s
are dened as jumps of f and its normal derivative
across S: m = 1fU and s = 1q
n
fU with n dened
as the unit normal vector pointing towards the ow.
Source distributions are responsible for displacing
the unperturbed ow so that the solid boundaries
are shaped as ow surfaces and therefore are dened
on S
B
. Dipoles are added so as to develop
circulation into the ow to simulate lift. They are
dened on S
W
and the part of S
B
referring to the
lifting components. In fact a surface distribution of
m is identied to minus the circulation around a
closed circuit which cuts the surface on which m is
dened at one point: G = m.
An important result given by Hess [36] states that
the ow induced by a dipole distribution m dened on
S
m
is given by a generalization of the BiotSavart law:
V
_
S
m
m
/
n
/
(x x
/
)
4p x x
/
[ [
3
dS
/
=
_
S
m
V
/
m n
/
( ) (x x
/
)
4p x x
/
[ [
3
dS
/

_
@S
m
m
/
s
/
(x x
/
)
4p x x
/
[ [
3
dS
/
(2:2:9)
where the line integral is taken along the boundary of
S
m
and s is the unit tangent vector to qS
m
in the
anticlockwise sense. If S
m
is a closed surface the line
integral vanishes, whereas if m is piecewise constant as
in the vortex lattice method, the surface term will
vanish leaving only the line term which corresponds
to a closed-loop vortex lament present along all lines
of m discontinuity on S
m
. The two terms on the RHS
of (2.2.9) have the form as the BiotSavart law
(2.2.5). From this analogy, c = Vm n is called
surface vorticity and ms line vorticity, which justies
the term vortex sheet for the wake of lifting bodies.
In potential theory a wake surface is the idealiza-
tion of a shear layer in the limit of vanishing
thickness. For an incompressible ow, the ow will
exhibit a velocity jump: 1VU]
W
= Vm
W
while
1VU]
W
n = 0, 1pU]
W
= 0. Using Bernoullis equa-
tion, it follows that
1pU]
W
r
=
qm
W
qt
V
W;m
1VU]
W
=
qm
W
qt
V
W;m
V
_ _
m
W
= 0, (2:2:10)
where V
W;m
= V

W
V

W
_ _
=2. Since, G = m
W
,
Kelvins theorem is obtained from (2.2.9) provided
that S
W
is a material surface moving with the mean
ARTICLE IN PRESS
n
r
: ,
B
S
:
W W
S
( ) P
F
ext
U
r
C

C
=
W
Fig. 4. Notations for the potential ow around a wing.
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 292
ow V
W;m
. Circulation will be materially conserved
and therefore m
W
is identied with its value at t = 0.
For a lifting problem, this means that m
W
is known
from the history of the wing loading. Assuming that
S
W
starts at the trailing edge of the wing, the
generation of the wake can be viewed as a
continuous release of vorticity in the free ow.
The streak line of a point along the trailing edge will
reveal the history of the loading of the specic wing
section as indicated in Fig. 7.
The rst model developed within the above
context is Prandtls lifting-line theory, see e.g. [38].
It concerns a lifting body of vanishing chord (or else
large aspect ratio), and thickness (Fig. 5). So s 0
whereas S
B
becomes a line carrying the loading G(y)
(bound vorticity), which is the only unknown since
the vorticity in the wake (trailing vorticity) is given
by @
y
G(y). In Prandtls original model, G(y) is
determined from airfoil data as equation (2.2.1).
Then, as an introduction to lifting-surface theory,
bound vorticity was placed along the c/4 line while
along the 3c/4 the non-penetration condition was
applied in order to determine G(y; t). The next step
was to introduce the lifting body as a lifting surface.
The most widely used model in this respect is the
vortex-lattice model [39]. It consisted of dividing S
B
and S
W
into panels and dening on them piecewise
constant m distributions (Fig. 6). Then according to
(2.2.9) the perturbation induced by the wing and its
wake, is generated by a set of closed-loop vortex
laments each dened along the boundary of a
panel. The dipole intensities on the wing can be
determined by the non-penetration condition at the
panel centres whereas along the trailing edge, see
(2.2.10) m = m
W
to ensure zero loading locally. In
the case of an unsteady ow, the loading G
B
will
change so that the vorticity shed in the wake will
also have a cross component qG
B
=qt:dt, as
indicated in Fig. 6.
Having determined m it is possible to calculate the
induced velocity and thus the angle of attack and
nally the loads from an airfoil data table look-up.
Another option frequently used in propeller appli-
cations, is to determine lift by integrating the
pressure jump along the section. Then by consider-
ing that the lift force is perpendicular to the effective
inow direction the angle of attack is determined. In
this case the pressure jump is obtained directly from
Bernoullis equation over the section except at the
leading edge where the geometrical singularity of a
blade with no thickness makes it necessary to
include the so-called suction force. In propeller
applications where this concept was rst introduced,
the suction force is estimated by means of semi-
empirical modelling [40]. Whenever used in wind
energy applications, the suction force has been
determined as rV Gdl, where V; G are the local
values at the leading edge and dl represents the
vector length along the leading edge line. In general,
the two schemes give comparable results. It is
difcult to clearly state which scheme is better to
use, since deviations appear as the angle of attack
increases so that a theoretical justication based on
matched asymptotic expansions is difcult. Another
point of concern, regarding both lifting theories, is
the detail in which the ow can be recorded. The
fact that the ow geometry is approximate suggests
that only at some distance from the solid boundary
the ow could be meaningful. Finally, for the same
reason, viscous corrections based on boundary layer
theory cannot be applied.
In order to overcome these difculties, the exact
geometry of the ow had to be included. This was
done by Hess who rst introduced the panel method
ARTICLE IN PRESS
= y .y
(y)
w
y

Fig. 5. The lifting-line model.
emission line
Zero loading at
trailing edge
S
e
S
e
W B
B


= .t

t
=
w

Fig. 6. The lifting-surface model.


M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 293
in its full form [41]. Now the panel grid is dened on
the true solid boundary and a piecewise constant
source distribution is introduced which can be
determined by the no-penetration boundary condi-
tion. In order to account for lift dipole distributions
are added. Because there is no kinematic condition
to determine m, Hess dened m to vary linearly along
the contour of each section (Fig. 7): m(s; y; t) =
s G(y; t)=L(y) and used (2.2.10) at the trailing edge
as an extra condition for determining G(y; t) (Fig.
7). The resulting problem is non-linear and an
iterative procedure must be used which penalizes the
computational cost considerably, as compared to
the lifting-surface model.
An important aspect of potential ow models
concerns wake dynamics. As discussed earlier, the
wake of a lifting body is a moving surface and S
W
should be allowed to change in time. Regarded as a
vortex sheet, the evolution of S
W
in time will be
subjected to convection and deformation. For
example in the case of the vortex lattice method,
each wake segment will conserve its intensity but its
vector length dl will satisfy the following equation:
d
dt
dl = (dl:V)V. (2.2.11)
As time evolves, wake deformations will generate
singularities such as intense roll-up along the wake
extremities and crossings. In fact at nite time the
ow will blow-up. In order to avoid blow-up,
corrective actions must be taken. If the simulation
retains the connectivity of the wake surface, the
lines dening the wake must be smoothened
regularly during the run time. Alternatively one
can apply remeshing which consists in dividing the
wake vortex segments so that they do not exceed a
prescribed upper limit. Both schemes, however,
require substantial bookkeeping and quite intense
procedures. With panel methods this problem
becomes more complicated because the wake will
also contain a surface vorticity term. A completely
different approach is to discard wake connectivity.
Rehbach [42] was the rst to note that for an
incompressible ow, vorticity concentrations dO of
the type o dD, g dS and Gdl behave similarly. Their
kinematic analogy was already discussed with
reference to (2.2.9). Dynamically they all satisfy
the same evolution equation:
D
Dt
d O
q
qt
d O (u:V)d O = (d O:V)u, (2.2.12)
where D=Dt denotes the total or material time
derivative. So Rehbach integrated the wake vorti-
city into point vortices, which subsequently moved
freely as uid particles carrying vorticity. This
procedure provides substantial exibility in the
evolution of the wake. He also introduced the
concept of modifying the kernel of the BiotSavart
law in order to cancel the r
2
singularity. The
theoretical justication of Rehbachs method came
later on leading to the development of the vortex
blob method [43].
In vortex models, the wake structure can either be
prescribed or computed as a part of the overall
solution procedure. In a prescribed vortex techni-
que, the position of the vortical elements is specied
from measurements or semi-empirical rules. This
makes the technique fast to use on a computer, but
limits its range of application to more or less well-
known steady ow situations. For unsteady ow
situations and complicated wake structures, free
wake analysis become necessary. A free wake
method is more straightforward to understand and
use, as the vortex elements are allowed to convect
and deform freely under the action of the velocity
eld as in Eq. (2.2.12). The advantage of the method
lies in its ability to calculate general ow cases, such
as yawed wake structures and dynamic inow. The
disadvantage, on the other hand, is that the method
is far more computing expensive than the prescribed
wake method, since the BiotSavart law has to be
evaluated for each time step taken. Furthermore,
free wake vortex methods tend to suffer from
stability problems owing to the intrinsic singularity
in induced velocities that appears when vortex
elements are approaching each other. To a certain
extent this problem can be remedied by introducing
a vortex core model in which a cut-off parameter
ARTICLE IN PRESS
L
= /L
W E
emission time
=
TE
t

TE
t t

2 TE t t


3
TE
t t


wake surface
streak line
( ; )=
B
TE
y t
e
m
i
s
s
i
o
n

l
i
n
e
s
T
Fig. 7. The exact potential model of a wing.
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 294
models the inner viscous part of the vortex lament.
In recent years, much effort in the development of
models for helicopter rotor ow elds have been
directed towards free-wake modelling using ad-
vanced pseudo-implicit relaxation schemes, in order
to improve numerical efciency and accuracy, e.g.
[44,45].
To analyse wakes of horizontal axis wind
turbines, prescribed wake models have been em-
ployed by e.g. [4648]. Free vortex modelling
techniques have been utilized by e.g. [49,50].
A special version of the free vortex wake methods
is the method described in [51], where the wake
modelling is taken care of by vortex particles or
vortex blobs.
Recently, the model of [52] was employed in the
NREL blind comparison exercise [137], and the
main conclusion from this was that the quality of
the input blade sectional aerodynamic data still
represents the most central issue to obtaining high-
quality predictions. Nevertheless, it is worth noti-
cing that by introducing relaxing techniques in the
wake evolution [53], it is nowadays possible to run a
large number of revolutions which is of importance
in aeroelasticity with reference to fatigue and
stability analysis; see later.
An alternative to panel methods is offered by the
Boundary Integral Equation Methods (BIEM). By
assuming stagnant ow inside the blade, m = f
and s = @
n
f, the resulting integral equation is only
weakly singular and so less expensive. Within the
eld of wind turbine aerodynamics, BIEMs have
been applied by e.g. [5456]; up to now, however,
only in simple ow situations.
Vortex methods have been applied on wind
turbine rotors particularly in order to better under-
stand wake dynamics. The next and quite challen-
ging step is to upgrade potential ow methods so as
to also include viscous effects. Examples of applying
viscousinviscid coupling within the context of 3D
boundary layer theory can be found in [26,57]. Also
attempts to include separation were made in [58].
All these works, however, cannot be considered
conclusive. There are several unresolved issues such
as convergence at the inboard region where
signicant radial ow develops as a result of
substantial separation, and the end conditions at
the tip. The fact that current trends in wind turbine
design indicate preference to pitch regulated ma-
chines could increase the interest in ow models
based on inviscid considerations. Finally, another
application of potential ow models is to use them
in order to obtain far eld conditions for RANS
computations in view of reducing their computa-
tional cost [59].
2.3. Generalized actuator disc models
The actuator disc model is probably the oldest
analytical tool for analysing rotor performance. In
this model, the rotor is represented by a permeable
disc that allows the ow to pass through the rotor,
at the same time as it is subject to the inuence of
the surface forces. The classical actuator disc
model is based on conservation of mass, momentum
and energy, and constitutes the main ingredient in
the 1D momentum theory, as originally formulated
by Rankine [60] and Froude [61]. Combining it with
a blade-element analysis, we end up with the
celebrated Blade-Element Momentum Technique
[6]. In its general form, however, the actuator disc
might as well be combined with the Euler or NS
equations. Thus, as will be shown in the following,
no physical restrictions have to be imposed on the
kinematics of the ow.
A pioneering work in analysing heavily loaded
propellers using a non-linear actuator disc model is
found in [62]. Although no actual calculations were
carried out, this work demonstrated the opportu-
nities for employing the actuator disc on compli-
cated congurations, such as ducted propellers and
propellers with nite hubs. Later improvements,
especially on the numerical treatment of the
equations, are due to [63,64]. Recently, Conway
[65,66] has developed further the analytical treat-
ment of the method. Within wind turbine aero-
dynamics [67] developed a semi-analytical actuator
cylinder model to describe the ow eld about a
vertical-axis wind turbine. A thorough review of
classical actuator disc models for rotors in general
and for wind turbines in particular can be found in
the dissertation [68]. Later developments of the
method have mainly been directed towards the use
of the NS or Euler equations.
In a numerical actuator disc model, the NS (or
Euler) equations are typically solved by a second-
order accurate nite difference/volume scheme, as in
a usual CFD computation. However, the geometry
of the blades and the viscous ow around the blades
are not resolved. Instead the swept surface of the
rotor is replaced by surface forces that act upon the
incoming ow. This can either be implemented at a
rate corresponding to the period-averaged mechan-
ical work that the rotor extracts from the ow or by
ARTICLE IN PRESS
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 295
using local instantaneous values of tabulated airfoil
data.
In the simple case of an actuator disc with
constant prescribed loading, various fundamental
studies can easily be carried out. Comparisons with
experiments have demonstrated that the method
works well for axisymmetric ow conditions and
can provide useful information regarding basic
assumptions underlying the momentum approach
[6972], turbulent wake states occurring for heavily
loaded rotors [73], and rotors subject to coning
[74,75].
In Fig. 8, an example of how various wake states
can be investigated by introducing a constantly
loaded actuator disc into the axisymmetric NS
equations is shown. By changing the thrust coef-
cient all types of ow states can be simulated,
ranging from the wind turbine state through the
chaotic wake state to the propeller state.
The generalized actuator disc method resembles
the BEM method in the sense that the aerodynamic
forces has to be determined from measured airfoil
characteristics, corrected for 3D effects, using a
blade-element approach. For airfoils subjected to
temporal variations of the angle of attack, the
dynamic response of the aerodynamic forces
changes the static aerofoil data and dynamic stall
models have to be included. However, corrections
for 3D and unsteady effects are the same for
generalized actuator disc models and the BEM
model, hence the description of how to derive
aerofoil data is the same as in Section 2.1.
In helicopter aerodynamics combined NS/actua-
tor disc models have been applied by e.g. [76] who
solved the ow about a helicopter employing a
chimera grid technique in which the rotor was
modelled as an actuator disk, and [77] who
modelled a helicopter rotor using time-averaged
momentum source terms in the momentum equa-
tions.
Computations of wind turbines employing nu-
merical actuator disc models in combination with a
blade-element approach have been carried out in
e.g. [69,70,78,79] in order to study unsteady
phenomena. Wakes from coned rotors have been
studied by Madsen and Rasmussen [74], Mikkelsen
et al. [75] and Masson et al. [78], rotors operating in
enclosures such as wind tunnels or solar chimneys
were computed by Hansen et al. [80], Phillips and
Schaffarczyk [81] and Mikkelsen and Srensen [82],
and approximate models for yaw have been
implemented by Mikkelsen and Srensen [83] and
Masson et al. [78]. Finally, techniques for employ-
ing the actuator disc model to study the wake
interaction in wind farms and the inuence of
thermal stratication in the atmospheric boundary
layer have been devised by Masson [84] and
Ammara et al. [85].
The main limitation of the axisymmetric assump-
tion is that the forces are distributed evenly along
ARTICLE IN PRESS
Fig. 8. Various wake states computed by actuator disc model with prescribed loading: (a) wind turbine state; (b) turbulent wake state; (c)
vortex ring state; (d) hover state. Reproduced from [70,73].
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 296
the actuator disc, hence the inuence of the blades is
taken as an integrated quantity in the azimuthal
direction. To overcome this limitation, an ex-
tended 3D actuator disc model has recently been
developed [86]. The model combines a 3D NS
solver with a technique in which body forces
are distributed radially along each of the rotor
blades. Thus, the kinematics of the wake is
determined by a full 3D NS simulation whereas
the inuence of the rotating blades on the ow
eld is included using tabulated airfoil data to
represent the loading on each blade. As in the
axisymmetric model, airfoil data and subsequent
loading are determined iteratively by computing
local angles of attack from the movement of
the blades and the local ow eld. The concept
enables one to study in detail the dynamics of
the wake and the tip vortices and their inuence
on the induced velocities in the rotor plane. A model
following the same idea has recently been sug-
gested by Leclerc and Masson [87]. A main
motivation for developing such types of model is
to be able to analyse and verify the validity of
the basic assumptions that are employed in the
simpler more practical engineering models. Re-
views of the basic modelling of actuator disc
and actuator line models can be found in [88], that
also includes various examples of computations.
Recently, another Ph.D. dissertation [89] carried out
a simulation employing more than four million
mesh points in order to study the structure of tip
vortices. In the following we will give some
examples of how the actuator disc/line technique
may help in understanding basic features of wind
turbine ows.
Computed iso-contours of vorticity for a three-
bladed rotor with airfoil characteristics corres-
ponding to the Tjreborg wind turbine is shown
in Fig. 9. In Fig. 10, a similar computation
shows the formation of the trailing tip vortices. It
is remarkable that the vortices are clearly visible
more than 3 turns downstream. A new and
interesting application of the actuator line model
is to study the interaction between two or more
turbines, especially for simulating park effects. In
Fig. 11, the outcome of a computation in which the
interaction between two wind turbines is simulated
by replacing the two rotors by actuator lines with
forces obtained from airfoil data is shown. Pre-
sently, this technique is used to investigate the effect
of large wind farms including many up- and
downstream wind turbines.
2.4. Navier Stokes solvers
2.4.1. Introduction to computational rotor
aerodynamics
The rst applications of CFD to wings and rotor
congurations were studied back in the late
seventies and early eighties in connection with
airplane wings and helicopter rotors [9094] using
ARTICLE IN PRESS
Fig. 9. Actuator line computation showing vorticity contours
and part of computational mesh around a three-bladed rotor.
Reproduced from [88].
Fig. 10. Iso-surface of constant vorticity showing the formation
of tip and root vortices. Reproduced from [89].
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 297
potential ow solvers. To overcome some of the
limitations of potential ow solvers, a shift towards
unsteady Euler solvers were seen through the
eighties [9598]. When computing power allowed
the solution of full Reynolds Averaged NS equa-
tions, the rst helicopter rotor computations in-
cluding viscous effects were published in the late
eighties and early nineties [99102].
In the late nineties, with the CFD solvers capable
of handling viscous ow around rotors, application
to wind turbine rotors became of practical interest.
The rst full NS computations of rotor aero-
dynamics was reported in the literature in the late
nineties [103107]. The European effort to apply NS
solvers to rotor aerodynamics had been made
possible through a series of National and European
project through the nineties. The European projects
dealing with development and application of the NS
method to wind turbine rotor ows was the Viscous
Effects on Wind turbine Blades (VISCWIND) from
1995 to 1997 [108], Viscous and Aeroelastic effects on
Wind Turbine Blades, (VISCEL), 1998 to 2000
[109,110], and Wind Turbine Blade Aerodynamics
and Aeroleasticity: Closing Knowledge Gaps, 2002 to
2004 [111115].
2.4.2. Approaches
As a consequence of the origin of most CFD
rotor codes from the aerospace industry and related
research, many existing codes are solving the
compressible NS equations and are intended for
high-speed aerodynamics in the subsonic and
transonic regime [116120]. For the helicopter
applications, where compressibility plays an impor-
tant role, this is the natural choice. For wind turbine
applications, however, the choice is not as obvious;
one reason being the very low Mach numbers near
the root of the rotor blades. As the ow here
approaches the incompressible limit, Mach0.01 it
is very difcult to solve the compressible ow
equations. One remedy to improve their capability
is the so-called preconditioning, that changes the
eigenvalues of the system of the compressible ow
equations by premultiplying the time derivatives by
a matrix. On the other hand, the compressible
solvers have many attractive features, among
these the ease of implementation of overlapping
and sliding meshes, application of high-order up-
wind schemes, and very well-developed solutions
methods.
Another very popular method, especially in the
US is the Articial Compressibility Method
[121,122], where an articial sound speed is intro-
duced to allow standard compressible solution
methods and schemes to be applied for incompres-
sible ows. In case of transient computations sub-
iterations are taken within each time step to enforce
incompressibility [122]. The method has several
attractive features: Among these a similar ease of
implementation of overlapping grids as the com-
pressible codes. Overlapping grids are a necessity, to
solve rotor/stator problems that are present when
the rotor, tower and nacelle are all included in the
computations. The main shortcoming of the method
may be problems to enforce incompressibility in
transient computations without the need for a huge
amount of sub-iterations, and the problem of
determining the optimum articial compressibility
parameter.
Due to the low Mach number encountered in
wind turbine aerodynamics, an obvious choice is
thus the incompressible NS equations. These
methods are generally based on treating pressure
as a primary variable [123125]. Extensions to
general curvilinear coordinates can be made along
the lines of [126]. The method is not as easily
extended to overlapping grids as the compressible
and the articial compressibility method, due to the
elliptical pressure correction equation. But the
method is well suited for solving the nearly
incompressible problems often experienced in con-
nection with wind energy. In connection with
steady-state problems, the method can be acceler-
ated using local time stepping, while the method
using global time stepping still is well suited for
transient computations.
ARTICLE IN PRESS
Fig. 11. Interaction of the wake between two partly aligned wind
turbines. Reproduced from [88].
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 298
2.4.3. Turbulence and transition
It is well known, that the NS equations cannot be
directly solved for any of the cases of practical
interest to wind turbines, and that some kind of
turbulence modelling are needed. The standard
approach to derive turbulence models is by time
averaging the NS equation, resulting in the so-called
Reynolds Averaged NS equations (RANS). Several
different models have been used with good results
for wind turbine applications, the most successful
ones being the k-omega SST model of Menter [127],
the SpalartAllmaras model [128], and the Bald-
winBarth model [129]. The BaldwinLomax [130]
model, often used in connection with helicopter and
x-wing applications, are not very well suited for
wind turbine applications, where relatively high
angles of attack are very common.
Several studies performed for stall controlled
wind turbines, have shown that all RANS models
lack the capability to model the stalled ow regime
at high wind speeds. One possible way around this
problem, the so-called Detached Eddy Simulation
(DES) technique [131,132], has shown some promis-
ing results but still needs further validation.
Additionally, the DES technique is much more
computationally expensive than the standard
RANS approach, as it needs much ner computa-
tional meshes and the computations needs to be
computed with time accurate algorithms.
From experiments it is known that laminar/
turbulent transition inuences the ow over rotor
blades for some cases. It has been demonstrated
for 2D applications, that transition models can
greatly improve the accuracy for cases where
transition phenomena are important. Even though
nearly all rotor studies so far have been com-
puted assuming fully turbulent conditions, it is
generally accepted that it is important to in-
clude laminar/turbulent transition to model the
physics as close as possible [133,134]. Predicting
transition in 3D is a much more complex task than
dealing with 2D, and 3D transition is an active
research eld.
2.4.4. Geometry and grid generation
To compute a rotor using CFD, the rst step is to
obtain a digitized description of the blade geometry.
Often the blade descriptions are given as spanwise
sectional information, listing the airfoil section, the
twist, the thickness, and the position with respect to
the blade axis. Often the blades are highly twisted,
and with a large taper in the spanwise direction.
Depending on the ow solver, different ap-
proaches to the mesh generation process exist:
Cartesian cut cells, unstructured and structured
and combinations of these. So far the majority of
ow solvers applied to wind turbine research have
utilized structured grids with hexahedral cells. In
connection with structured grids there are several
issues that need to be decided upon. Generally, the
problem of making a high quality grid around a
modern rotor cannot be handled by a single block
conguration, but needs some kind of multi block
mesh. These can either be conforming at the block
boundaries, non-conforming or overlapping. The
overlapping grids gives the highest degrees of
freedom followed by the non-conforming and the
conforming grids. Firstly, the grid needs to accu-
rately resolve the blade shape, with good resolution
of the leading edge and tip region. Secondly, the
grids also need to resolve the regions around the
blade with sufcient resolutions to capture the ow
physics. As the Reynolds numbers are quite high,
16 million, the cells near the rotor blades become
very thin, as the non-dimensional distance y+ must
be approximately 1 to resolve the laminar sub-layer
and have accurate solutions. The mesh generation
process calls for some degree of experience and grid
renement studies to verify that the grid is sufcient
to resolve the desired physics. Also the grids need to
extend far away from the rotor, in the order of
several rotor diameters, to avoid disturbing the
induced velocity eld near the rotor blades. For
axial ow conditions, the ow solvers often take
advantage of the rotational periodicity of the rotor,
solving only for a single blade using periodic
conditions.
Using an unstructured ow solver with tetrahe-
dral cells, the grid generation process is less
cumbersome. But the problem of resolving very
thin boundary layers using tetrahedral cells is well
known, and it may be necessary to combine the
solver with some kind of prismatic grids near the
blade surface to avoid this problem. The use of
unstructured ow solvers is not wide spread in
connection with wind turbine aerodynamics, prob-
ably because of the limited geometrical complexity,
and the strength of unstructured solvers mainly
being their ability to cope with complex geometries.
2.4.5. Numerical issues
The codes typically used for wind turbines are of
at least second-order accuracy in both time and
space, often with an implicit time discretization
ARTICLE IN PRESS
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 299
scheme to loosen the time step restriction inherent
to explicit methods. Typically, the viscous terms are
discretized with central differences, while the con-
vective terms are discretized with second- or third-
order upwind schemes. To solve routinely for 510
million grid points, the solvers are often available in
a parallelized version that allows for execution on
several CPUs in parallel. The rotating nature of the
problem requires the use of either a moving frame
including the non-inertial acceleration terms or a
moving mesh option where so-called mesh uxes
must be included in the code. For a good overview
of the numerical issues in connection with incom-
pressible ow, see [135].
2.4.6. Application of CFD to wind turbine
aerodynamics
The major part of wind turbine rotor computa-
tions performed until now has been focused on zero
yaw rotor only conguration, where the nacelle and
tower have been neglected, and the inow to the
rotor has been assumed to be steady without shear.
This is, of course, a great simplication, but in many
cases still a sufciently good approximation. The
effect of the tower on the rotor on an upwind
turbine is comparable to other unsteady effects,
such as incoming turbulence, time variations of the
rotor and of the incoming ow. A simulation,
working with a full turbine geometry has been tried
[106]. This type of simulation is much more
expensive, and needs some kind of sliding/over-
lapping mesh to accommodate the movement of the
rotor with respect to the turbine tower and nacelle.
Additionally, the simulation needs to be time
accurate, and good resolution of the ow around
the tower is needed to capture the tower wake far
downstream of the turbine.
One of the rst real proofs that CFD for wind
turbine rotor applications can be useful came in
connection with the blind comparison organized by
the National Renewable Energy Laboratory in
Boulder, Colorado in December 2000 [136138].
Some of these results were later published in
[139,140]. Here several wind turbine research groups
were asked to compute a series of different
operational conditions for the NREL Phase-VI
turbine, corresponding to actual cases measured in
the NASA Ames 80 120 ft wind tunnel. When the
results were made publicly available, it proved that
one of the applied CFD codes were consistently
reproducing the measured distribution of the aero-
dynamic forces along the blade span, even under
highly 3D and extreme stall conditions.
The output extracted from typical CFD rotor
computations are the low-speed shaft torque, or
power production, and root ap moments, see Fig.
12. For modern full size commercial turbines, the
power and root ap moment are typically the only
available properties. Besides the quantities normally
measured, the CFD simulations provide a huge
amount of detailed information that can be used to
provide more insight. The data typically extracted
are the spanwise distributions of force coefcients,
Fig. 13, the limiting streamlines on the blade
surfaces, Fig. 14, and the sectional pressure
distributions along the blade span, Fig. 15. For
the NREL Phase-VI turbine, these detailed quan-
tities can be compared to measurements, but this is
not generally the case for commercially available
turbines.
Another application of rotor CFD is the study of
different aerodynamic details of the rotor, such as
the blade tips, the design of the root section etc.
Here, the CFD technique can be used to supply
ARTICLE IN PRESS
700
800
900
1000
1100
1200
1300
1400
1500
1600
1700
6 8 10 12 14 16 18 20 22 24 26
L
o
w

S
p
e
e
d

S
h
a
f
t

T
o
r
q
u
e

[
N
m
]
Wind Speed [m/s]
Measured
Computed
1000
1500
2000
2500
3000
3500
4000
4500
5000
6 8 10 12 14 16 18 20 22 24 26
R
o
o
t

F
l
a
p

M
o
m
e
n
t

[
N
m
]
Wind Speed [m/s]
Measured
Computed
Fig. 12. Comparison of computed and measured low-speed-shaft torque and root ap moment for the NREL Phase-VI rotor during axial
ow conditions. The 7 one standard deviation of the measurements is indicated around the measured values.
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 300
information, that the engineering methods are not
capable of providing.
With the increase in computational power, it is
today both possible and affordable to do yaw
computations using CFD [133,141143]. In contrast
to the axial ow cases, the total rotor must be
modelled in yaw simulations, thereby increasing the
number of mesh points typically by a factor of
three. Additionally, the azimuthally variation in-
herent in yaw simulations dictates a time accurate
simulation, as no steady state solution can exist,
thereby increasing the computing time severely.
Typically, a yaw computation will be 1020 times
more expensive compared to steady state axial ow
computations. Fig. 16 shows the normal and
tangential force at the r/R = 30% radius during
one revolution at 601 yaw operation.
The release of the measurements on the NREL
Phase-VI rotor has heavily inuenced the CFD
activities dealing with wind turbine rotor aerody-
namics. This unique data set, with several well-
documented cases, has given a new possibility to test
details of state of the art CFD codes. In the years
since the release of the measurements nearly half of
all published CFD studies of wind turbine rotors
deals with these measurements. Besides the refer-
ences mentioned other places in this paper, the
following studies use the NREL/NASA Ames
measurements [144147].
Recently DES simulations of rotors at realistic
operational conditions have been attempted [148].
Again the NREL Phase-VI rotor was used to verify
the model. The reason for this choice is, besides the
availability of detailed measurements, the fact that
the rotor has a limited aspect ratio, hence making
the DES computations more affordable with respect
to the number of grid cells. The mesh for this
simulation consists of 15 million cells, compared to
around 2 million cells for a standard RANS rotor
computation. The fact that DES computations must
be performed using time accurate computations,
makes these types of simulations 2040 times more
ARTICLE IN PRESS
0
0.5
1
1.5
2
2.5
3
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
C
N
r/ R
Measured
Computed
-0.08
-0.06
-0.04
-0.02
0
0.02
0.04
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
C
T
r/ R
Measured
Computed
Fig. 13. Spanwise normal and tangential force coefcients for run S20000000, of the NREL/NASA Ames measurements, corresponding
to a wind speed of 20 m/s.
NREL PHASE-6, W=10 [m/ s], Limiting Streamlines
RISOE, EllipSys 3D Computation
Fig. 14. Limiting streamlines on the surface of the NREL Phase-
VI rotor for the 10 m/s axial case. The picture shows a sudden
leading edge separation around r/R = 0.47.
-1.5
-1
-0.5
0
0.5
1
1.5
2
2.5
3
0 0.2 0.4 0.6 0.8 1
-
C
p
X/ Chord
r/ R=0.63
Fig. 15. Comparison of the measured and computed pressure
distribution at the r/R = 0.63 section of the NREL Phase-VI
blade, the measurements are shown as circles while the
computations are shown with lines.
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 301
expensive than standard steady-state rotor compu-
tations. For modern-type rotors with large aspect
ratios, the cell count would be even higher and the
computations even more expensive. For the NREL
Phase-VI rotor the improvement using the DES
technique is very limited, but for other rotors where
the RANS equations do not perform as well, DES
may provide much more improvement. The use of
pure Large Eddy Simulation, has been demon-
strated in [149], where a computational grid of 300
million points is used to compute the initial
transient of the development of a tip vortex.
2.4.7. Future
Today NS solvers are an important tool for
analysing different wind turbine rotor congura-
tions, and are used routinely along with measure-
ments and other computational tools for
development and investigation of wind turbines.
NS solvers are especially well suited for detailed
investigation of phenomena that cannot directly be
accessed by simpler and less computationally
expensive methods. Additionally NS methods can
be used as a supplement to measurements, where
they can be used both in the planning phase and to
have a better interpretation of the actual physics in
connection with the analysis of measurements.
Finally, one of the latest trends is to couple NS
solvers to structural codes to perform full elastic
computations of wind turbine rotors.
3. Structural modelling of a wind turbine
The main purpose of a structural model of a wind
turbine is to be able to determine the temporal
variation of the material loads in the various
components. This is accomplished by calculating
the dynamic response of the entire construction
subject to the time-dependent load using an aero-
dynamic model, such as the BEM method. For
offshore wind turbines also wave loads and perhaps
ice loads on the bottom of the tower must be
estimated. Two different and frequently used
approaches to set up a dynamic structural model
for a wind turbine are described in the next
subsections.
3.1. Principle of virtual work and use of modal shape
functions
The principle of virtual work is a method to set
up the correct mass matrix, M, stiffness matrix, K,
and damping matrix, C, for a discretized mechanical
system as
M x C_ x Kx = F
g
, (3.1.1)
where F
g
, denotes the generalized force vector
associated with the external loads, p. Eq. (3.1.1) is
of course nothing but Newtons second law assum-
ing linear stiffness and damping, and the method of
virtual work is nothing but a method that helps
setting this up for a multibody system and that is
especially suited for a chain system. Knowing the
loads and appropriate conditions for the velocities
and the deformations, Eq. (3.1.1) can be solved for
the accelerations wherefrom the velocities and
deformations can be determined for the next time
step. The number of elements in, x, is called the
number of DOF, and the higher this number the
more computational time is needed in each time step
to solve the matrix system. Use of modal shape
functions is a tool to reduce the number of DOF
and thus reduce the size of the matrices to make the
computations faster per time step. A deection
ARTICLE IN PRESS
-2
0
2
4
6
8
10
12
0 50 100 150 200 250 300 350
C
N

Azimuth Angle [deg.]
S1500600, r / R=0.30
Measurements
Computed
Measurements
Computed
-0.4
-0.2
0
0.2
0.4
0.6
0.8
1
0 50 100 150 200 250 300 350
C
T

Azimuth Angle [deg.]
S1500600, r/ R=0.30
Fig. 16. Azimuth variation of normal and tangential force coefcient, for the NREL Phase-VI turbine at the r/R = 0.4 section during a
601 yaw error at 15 m/s wind speed.
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 302
shape is here described as a linear combination of a
few but physical realistic basis functions, which are
often the deection shapes corresponding to the
lowest eigenfrequencies (eigenmodes). For a wind
turbine such an approach is suited to describe the
deection of the tower and the rotor blades and the
assumption is that the combination of the Power
Spectral Density of the loads and the damping of
the system do not excite the eigenmodes associated
with higher frequencies. In the commercially avail-
able and widely used aeroelastic simulation tool,
FLEX, see e.g. [150], only the rst 3 or 4 (2 apwise
and 1 or 2 edgewise) eigenmodes are used for the
blades. Results from this model are generally in
good agreement with measurements, indicating the
validity of the underlying assumption. First one has
to decide on the DOF necessary to describe a
realistic deformation of a wind turbine. For instance
in FLEX4 1720 DOFs are used for a three bladed
wind turbine, with 34 DOFs per blade as described
above, 4 DOFs for the deformation of the shaft (1
for torsion, 2 for the hinges just before the rst
bearing, with associated angular stiffness to describe
bending, and 1 for pure rotation), 1 DOF to
describe the tilt stiffness of the nacelle and, nally,
3 DOFs for the tower (1 for torsion, 1 for the rst
eigenmode in the direction of the rotor normal and
1 in the lateral direction).
The method of virtual work will only be briey
described. For a more rigorous explanation of the
method the reader is referred to textbooks on
dynamics of structures. The values in the vector
describing the deformation of the construction, x
i
,
are denoted the general coordinates. To each
generalized coordinate is associated a deection
shape, u
i
, that describes the deformation of the
construction when only x
i
is different from zero and
typically has a unit value. The element i in the
generalized force corresponding to a small displace-
ment in DOF number i, dx
i
, is calculated such that
the work done by the generalized force equals the
work done on the construction by the external loads
on the associated deection shape.
F
g;i
dx
i
=
_
S
p u
i
dS, (3.1.2)
where S denotes the entire system. Please note that
the generalized force can be a moment and that the
displacement can be angular. All loads must be
included, i.e. also gravity and inertial loads such as
Coriolis, centrifugal and gyroscopic loads. The non-
linear centrifugal stiffening can be modelled as
equivalent loads calculated from the local centrifu-
gal force and the actual deection shape as shown in
[151]. The elements in the mass matrix, m
i,j
, can be
evaluated as the generalized force from the inertia
loads from an unit acceleration of DOF j for a unit
displacement of DOF i. The elements in the stiffness
matrix, k
i,j
, correspond to the generalized force from
an external force eld which keeps the system in
equilibrium for a unit displacement in DOF j and
which then is displaced x
i
= d
ij
, where d
ij
is
Kroneckers delta. The elements in the damping
matrix can be found similarly. For a chain system
the method of virtual work as described here
normally gives a full mass matrix and diagonal
matrices for the stiffness and damping. For one
blade rigidly clamped at the root (cantilever beam)
it is relatively easy to estimate the lowest eigen-
modes (rst apwise u
1f
(x), rst edgewise u
1e
(x) and
second apwise u
2f
(x)) e.g. using an iterative method
as described in [151]. The eigenmodes are normally
described in a coordinate system aligned with the tip
chord as e.g. shown in Fig. 1. It is practical to
normalize the deection shapes so that the tip
deection is unity. It is now assumed that any
deection can be described as a linear combination
of these modes as
u(x) = x
1
u
1f
(x) x
2
u
1e
(x) x
3
u
2f
(x). (3.1.3)
The velocity and accelerations can be calculated,
respectively, as
_ u(x) = _ x
1
u
1f
(x) _ x
2
u
1e
(x) _ x
3
u
2f
(x) (3.1.4)
and
u(x) = x
1
u
1f
(x) x
2
u
1e
(x) x
3
u
2f
(x). (3.1.5)
The advantage of the method using generalized
coordinates and modal shape functions is that the
number of DOF in the dynamic system can be
reduced to a relatively small number. Further, some
high eigenfrequencies are ltered away, which is
benecial for the allowable time step, when comput-
ing deformations, x
n1
i
, and velocities, _ x
n1
i
, at time
t = (n+1)Dt from deformations, x
n
i
, velocities, _ x
n
i
,
and accelerations, x
n
i
, at time t = nDt. A good
choice for the time integration scheme is the
RungeKuttaNystro m method, which requires
for stability reasons that the time step should
resolve the highest eigenfrequency with 4 points;
but for accuracy reasons 10 points is preferred.
By reducing the highest eigenfrequency using a
modal description of e.g. the blades not only
ARTICLE IN PRESS
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 303
reduces the number of DOFs but also larger time
steps can be taken.
3.2. FEM modelling of wind turbine components
applying non-linear beam theory
Even though modal analysis offers a computa-
tionally effective way to analyse wind turbine
dynamics, most of the recently developed aeroelastic
codes use a full FEM approach [152], which allows
a more complex deformation state of the wind
turbine. The main features of such modelling
procedures together with some aspects of non-linear
beam theory are discussed in the present section.
The structural modelling of wind turbines is
based on beam theory. For a 1D structure (beam)
subjected to bending in two directions, torsion and
axial tension, the formulation of the problem
consists of two parts: (a) the elastic model which
reduces the 3D structure of each component into a
1D structure concentrated along the elastic axis of
the beam, and (b) the derivation of the dynamic
equations. In classical beam theory (rst order) the
deformed position r of any point initially at x
0
=
(x; y; z)
T
(Fig. 17) can be described by the displace-
ment vector u = {u; v; w; y]
T
consisted of the two
bending displacements u, w the torsion angle y and
the tension v:
r = n
0
S
0
u S
1
@
y
u;
S
0
=
1 0 0 z
0 1 0 0
0 0 1 x
_

_
_

_;
S
1
=
0 0 0 0
x 0 z 0
0 0 0 0
_

_
_

_:
(3.2.1)
Using Hookes law and assuming that shear stresses
do not produce net torsion, the sectional elastic
loads can be derived:
F
y
= (EA)q
y
v (EAx
+
)q
2
yy
u (EAz
+
)q
2
yy
w,
M
y
= (GI
t
)q
y
y,
M
x
= (EI
xz
)q
2
yy
u (EI
xx
)q
2
yy
w (EAz
+
)q
y
v,
M
z
= (EI
zz
)q
2
yy
u (EI
xz
)q
2
yy
w (EAx
+
)q
y
v,
(3:2:2)
where the terms in parentheses denote the averaged
sectional properties of the structure. By considering
the balance of loads and moments on of a beam
ARTICLE IN PRESS
Undeformed geometry
Deformed geometry
dy
y
x
z
A
u
w
u+du
w+dw r
Deformed elastic axis
M + dM
z
z
z
M + dM
y y
M + dM
x x
z
F + dF
y y
F + dF
x x
F + dF
Mx
F
x
M
z
F
z M
y
F
y
dr
r
a
P
x
z
y
A
(a)
(b)
a
b
Fig. 17. Kinematics and dynamics of a beam structure.
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 304
element of length dy the beam equations are
derived:
_
A
r dm
_ _
dy = dF
_
A
g dm
_ _
dy dp dy,
(3.2.3)
_
A
r
0
r dm
_ _
dy = dM dr (F dF)

_
A
r
0
g dm
_ _
dy r
a
dp dy; (3:2:4)
where m denotes the mass per unit length, dF, dM
are the net elastic loads on dy, g is the acceleration
of gravity and dp the sectional aerodynamic loads
exerted at the aerodynamic centre r
a
= (x
a
; 0; z
a
)
while
r
0
= {du x zy; dv dy; dw z xy]
T
{x zy; dy; z xy]
T
.
In view of a more systematic and general framework
for formulating dynamic equations, Hamiltons
principle has been also used [153,154].
By introducing (3.2.2) into (3.2.3) and (3.2.4) the
beam dynamic equations are obtained:
_
A
(S
0
)
T
r dm q
y
_
A
(S
1
)
T
r dm = q
y
[K
11
q
y
u]
q
2
yy
[K
22
q
2
yy
u] q
y
[K
12
q
2
yy
u]
q
2
yy
[K
21
q
y
u]
_
A
(S
0
)
T
g dm S
a
dP;
(3:2:5)
K
11
=
F
y
0 0 0
0 EA 0 0
0 0 F
y
0
0 0 0 GI
t
_

_
_

_
,
K
22
=
EI
zz
0 EI
xz
0
0 0 0 0
EI
xz
0 EI
xx
0
0 0 0 0
_

_
_

_
,
S
a
=
1 0 0
0 1 0
0 0 1
z
a
0 x
a
_

_
_

_
,
K
12
=
0 0 0 0
EAx
+
0 EAz
+
0
0 0 0 0
0 0 0 0
_

_
_

_
,
K
21
=
0 EAx
+
0 0
0 0 0 0
0 EAz
+
0 0
0 0 0 0
_

_
_

_
.
In case of exible blades if large deformations are
expected, as in the case of helicopter blades, second-
order non-linear beam theory is to be used. The
relevant developments originate from the work in
[154]. The beam axis again lies along the y-axis but
in the undeformed state of the beam while x and z
denote the two bending directions of the beam. At
its deformed state a local co-ordinate system O
/
xZz
is introduced which follows the pre-twist of the
beam so that x and z coincide with the local
structural principle axes of the cross section
(Fig. 18).
Then, (3.2.1) becomes
r =
u
y v
w
_

_
_

_
E
x
l@
y
y
z
_

_
_

_
, (3.2.6)
where l denotes the warping function of the cross
section and E is the transformation matrix between
Oxyz and O
/
xZz. In [154], E is given in terms of the
elastic displacements up to second order. Next the
strain tensor
ij
dened through (3.2.6) is introduced
in Hookes law in order to dene the straindispla-
cements relations, which are used in the derivation
of the resulting sectional elastic loads as in (3.2.2).
Because they are derived in the O
/
xZz system, before
introducing them in (3.2.3) and (3.2.4) they are
transformed into the Oxyz system using the
ARTICLE IN PRESS
x
z
y
u
w
v
O
O
z
x
z
x
u
w

+
Fig. 18. Beam co-ordinate system denition.
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 305
transformation matrix E. The nal expressions are
quite lengthy and the reader is referred to [154]. It is
noted that compared to the classical beam model,
second-order theory will produce fully complete
stiffness matrices containing many non-linear terms.
The above non-linear beam model is a specic
example amongst several models of varying com-
plexity which have been gradually developed start-
ing from the Timoshenko beam theory, by
eliminating the ad hoc simplifying assumptions of
the original model [155158].
Regarding wind turbines, almost all structural
models are based on classical beam theory
[150,153,159162]. This is partially due to the fact
that wind turbine components are far more rigid
compared to helicopter blades for which non-linear
theory has been developed. Furthermore most of the
difculties in analysing wind turbine systems are
generated by the aerodynamics and in particular the
onset of stall, which is certainly less pronounced on
helicopter rotors. Current designs are quite stiff so
that non-linear beam modelling is not expected to
change drastically the quality of predictions besides
providing a sound basis for including the geometrical
non-linearities [163]. However, if wind turbines
become more exible, it could become necessary to
adopt such kind of structural modelling.
Regardless the details the beam equations are
fourth order with respect to bending and second
order with respect to tension and torsion. So for a
FE approximation of the equations, C
1
shape
functions are used for u, w and C
0
for v and y. At
the element level, u, w are approximated with third-
order polynomials and the DOFs are the values and
space derivatives at the end nodes while v and y are
approximated with rst-order polynomials and the
DOFs again at the end nodes. In some models and
certainly when the second-order beam theory is
applied, second- and third-order polynomials are
used, respectively, for v and y. In this case the mid-
point as well the points at 1/3 and 2/3 interior points
are used as nodes. So within an element e,
u
e
(y; t) = N
e
(y) ^ u
e
(t); (3.2.7)
where N
e
(y) is the matrix containing the shape
functions and ^ u
e
(t) the vector of the DOFs at the
nodes of the elements [164]. As in Section 3.1
concerning the principle of virtual work, which in
the FEM terminology is the Galerkin formulation
of the problem is used to generate the discrete
equations. With reference to (3.2.5) taken symboli-
cally as Z(u; _ u; u) = 0, for any admissible virtual
displacement eld du we require that
_
L
0
du
T
Z(u; _ u; u) dy

e
_
L
e
0
du
T
e
Z(u
e
; _ u
e
; u
e
) dy = 0; \du. (3:2:8)
Because du
e
(y; t) = N
e
(y) d^ u
e
(t), it follows that:

e
_
L
e
0
N
eT
Z(N
e
^ u
e
; N
e
^
_ u
e
; N
e
^
u
e
) dy = 0, (3.2.9)
which is a set of second-order ordinary equations in
time with respect to ^ u
e
(t). Although Z can contain
non-linear terms, it can always be written in the
usual form:
M(u

)
^
u C(u

)
^
_ u K(u

) ^ u = Q(u

); (3.2.10)
where the mass, damping and stiffness matrices as
well as the generalized loads in the RHS, in general
will depend on the displacement eld and its time
and space derivatives (noted by a tilde).
ARTICLE IN PRESS

k
r
Gk
O
G
,O
x
G
, x
z
G
, z
y
G
, y
y
z
x
O
x
y
z
O
Fig. 19. Multi-body representation of a wind turbine.
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 306
The main components of a wind turbine are the
blades, the drive-train and the tower (Fig. 19). They
are all modelled as beam structures and typically the
structural properties are assumed for each compo-
nent to continuously vary along the corresponding
elastic axis. However, localized properties can be
added in the form of concentrated masses, dampers
or springs. The gearbox (if present), the generator,
the hub are usually added in this way. Other
examples are the exibility or damping character-
istics of the yaw bearing or the pitch mechanism.
The involvement of different body motions for each
component in combination with the connections
where loads and displacements are communicated
from one component to the other, calls for a global
formulation of the dynamic problem. To this end
most works adopt a multi-body approach [165],
which consists of considering each component
separately subjected to appropriate boundary con-
ditions, which t the different components into the
complete conguration. In such an approach the
elastic DOF of each component are dened as local
elastic displacements to which rigid body motions
are added through the kinematic boundary condi-
tions, see e.g. [161,166,167] for a more general
discussion. It is also possible to introduce the global
displacements and rotations at each node instead of
the local ones [153], allowing a unied description of
the complete conguration. In this case, however,
the non-linearities of the connections are introduced
implicitly and therefore linearization can only be
performed globally, which is not always desirable.
In multi-body modelling each component k is
assigned a local coordinate system Oxyz with its
undeformed elastic axis along the y-axis. Then the
position of a point with respect to the xed system
r
G;k
can be expressed with respect to its local
position r
k
(dened as in (3.2.1)) as follows:
r
G;k
= q
k
A
k
r
k
, (3.2.11)
where q
k
denotes the position of the origin of the
local system with respect to the xed system and A
k
denotes the local to global transformation matrix.
The exact form of q
k
and A
k
depends on the
kinematic conditions introduced when joining (con-
necting) the various components (i.e. blades-to-hub
or drive-train-to-tower). Depending on the type of
connection, common global displacements and
rotations are assigned for the restricted DOF, which
are identied to either an existing elastic DOF (shaft
bending at the hub) or a rigid body motion (blade
pitch). The component contributing the kinematics
will in response receive from the rest of the
connected components their internal (or reaction)
loads. This operation involves co-ordinate system
transformations between the components. The set
of kinematical DOF involved in the denition of q
k
and A
k
for all components is denoted collectively as
q. So q
k
= q
k
(q; t) and A
k
= A
k
(q; t). q
k
is dened as
a mixed series of translations and rotations, whereas
A
k
is dened solely as a sequence of rotations. A
typical example is shown in Fig. 20, where q
1
q
6
are
the elastic DOF at the top of the tower, q
8
q
13
are
the elastic DOF of the drive train at the hub and q
7
and q
14
are the yaw angle and the pitch of the blade,
respectively. By introducing (3.2.11) into (3.2.5), the
centrifugal and Coriolis terms of the inertia loads
appear as a result of the time derivatives of A
k
while
in the Hamiltons principle approach they are
produced automatically.
By combining the equations of all components
the complete system of the dynamic equations for
the wind turbine is obtained in the form of (3.2.10)
with respect to an extended vector of DOF
^ u
ext
= ^ u; q
_ _
. By construction q
k
and A
k
will depend
on unknown DOF while at the same time they
correlate the state of the components in contact so
the terms they generate are non-linear with respect
to the unknown DOF u
k
and q. Linearization in this
case is a tedious procedure which must be done
carefully and systematically. Fortunately software
using symbolic mathematics, such as Mathematica,
ARTICLE IN PRESS
y
O
q
8
, q
9
, q
10
, q
11
,
q
12
, q
13
y
G
, y
z
x
O
x
y
z
O
G
,O
x
G
, x
z
G
, z
q
1
, q
2
, q
3
, q
4
,
q
5
, q
6
q
7
q
14
Fig. 20. A typical denition of the q DOF on a HAWT.
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 307
are available which are strongly recommended. The
procedure is based on Taylors expansions of all
variables and retaining of rst-order terms. Alter-
natively, the extended version of (3.2.10) can be
solved directly through an iterative procedure in
each time step, which can use the linearized solution
as a starting predictor.
4. Problems and solutions in wind turbine
aeroelasticity
In the following section, the use of aeroelastic
codes on wind turbine constructions including
stability studies will be illuminated mainly through
examples.
4.1. Aeroelastic stability
Wind turbines suffer from low structural damp-
ing which can become critical under certain opera-
tional conditions. Most problems appear on the
blades that receive almost 100% of the loading. In
particular the onset of stall plays a decisive role, not
only on stall-regulated machines, but also on pitch-
regulated ones around rated conditions. On the
other hand, the blades are made of composite
materials for which the knowledge on damping is
substantially inferior compared to steel or concrete
constructions. The damping of composite structures
depends on the ambient temperature. So depending
on the season, or the time during the day, the
structural damping of the blades can decrease
substantially. Furthermore, aging always degrades
the damping of composite structures. There is a
denite need to increase structural damping and
signicant effort has been put recently on its
modelling in composite structures as well as in their
appropriate design. This is one direction of research
regarding the improvement of the stability char-
acteristics of modern wind turbines; see [168] for a
review of the current developments. Damping is
required to suppress the onset of vibrations
generated by the unsteady aerodynamic loads that
interact with the wind turbine structure. First, the
aeroelastic modelling of wind turbines is considered
and then on this basis the aeroelastic stability
problem is formulated and the relevant available
methods are reviewed.
Aerodynamic modelling involves the calculation
of the aerodynamic loads that also depend on the
dynamics of the system. The ap and lag vibrations
are interpreted by the incoming ow as a change in
the effective angle of attack indicating strong
aeroelastic coupling. All the currently available
aeroelastic tools for wind turbines use 2D dynamic
airfoil data models such as the BeddoesLeishmann
or the ONERA model. They provide aerodynamic
coefcients as functions of the aerodynamic state
variables that satisfy appropriate dynamic equa-
tions [17,169]. In either model the equations are 2D
and so they are applied on a strip-by-strip basis,
meaning that there is no interaction in the radial
direction. This ts well with BEM modelling of the
overall aerodynamic analysis. In fact the following
presentation is referred to the BEM context, since
most existing stability models are based on it. If a
potential ow model is applied, then the attached
part of lift can be associated to potential lift while
the rest is simply added [163]. If the potential ow
model is enhanced with a separation model there
should be correlation on both aspects [58]. How-
ever, the latter has not been applied yet in
aeroelastic computations.
By considering a blade section at a radial position
as shown in Fig. 1 the local aerodynamic loads
acting on this section of the blade can be written as
follows, since the lift is perpendicular and the drag
parallel to the relative wind:
p
y
= L sin f D cos f,
p
z
= L cos f D sin f,
M
y
= M, (4:1:1)
where L, D are the lift and drag forces, M is the
pitching moment, f is the local ow angle with
respect to the rotor plane. Once the induced velocity
has been determined, the local effective incidence a
and relative ow velocity can be calculated as shown
in Section 2.1. They will depend on the elastic
deformations and the velocities, the rigid body
motions and, if appropriate, orientation changes, as
in pitch and yaw variations. In order to obtain a
unied description of the coupled aeroelastic
problem a so-called aeroelastic nite element is
dened [170] which, in addition to the elastic DOF,
the aerodynamic state variables added. For example
in the ONERA model there will be in total eight
additional DOF. Because the dynamic equations for
the aerodynamic state variables only can be rst
order in time, in order to produce a consistent
coupled system, they can be differentiated as in
[171]. Therefore they can be assembled together
with the dynamic equations in the form of (3.2.10)
where the vector of the unknowns is further
ARTICLE IN PRESS
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 308
extended to include the additional aerodynamic
DOF. Another approach used in [172] is to
introduce the modal expansion of the structural
DOF at stand still, which better conditions the
coupled system.
On the above basis it is possible to formulate the
aeroelastic stability problem for wind turbines.
Within the context of linear theory, stability
boundaries are obtained by examining the evo-
lution of small perturbations with respect to a
steady state or a periodic solution acting as
reference. To this end the original equations are
linearized with respect to the reference state and
the resulting system solved with respect to the
uctuations:
M(u
ref
)
^
u C(u
ref
)
^
_ u K(u
ref
) ^ u = Q; (4.1.2)
where M, C and K depend only on the reference
solution. If the matrices do not depend on time,
(4.1.2) is reformulated into a rst-order system:
_ y = D y b; y = ^ u;
^
_ u
_ _
T
. (4.1.3)
The eigenvalue analysis of D provides the eigen-
frequencies and damping of the system [173]. This is
possible when the reference state corresponds to a
steady solution as in the case of an isolated blade.
However, when analysing the complete wind
turbine conguration, due to the rotation of the
blades, the reference state should correspond to a
periodic equilibrium state with reference to the rotor
speed, which is assumed constant. In this case the
corresponding theoretical framework is Floquets
theory, which for a large system is computationally
heavy [174]. If the blades are identical and the
number of blades NX3, which is the most frequent
case, it is possible by means of multi-blade
transformation to eliminate the periodic coefcients
in the coefcients of (4.1.2) and therefore be able
to still use eigenvalue analysis [175]. In this context
the non-linear equations of the system are inte-
grated in time until such a periodic state is attained.
In the case of an unstable situation, because the
time domain response will contain signicant
components in all of the basic eigenfrequencies of
the system, the reference state is obtained by
truncating the response so as to retain only its 1
and N/rev components. To this end all DOF in the
rotating system q
m
; m = 1; . . . NX3, are reformed:
q
m
= q
0
q
c
cos c
m
q
s
sin c
m
, where c
m
= Ot
(2p=N)(m 1) is the corresponding azimuth loca-
tion. Next the equations for the blade DOFs are
rearranged by applying the operators (1=N)

N
m=1
( ), (2=N)

N
m=1
( ) cos c
m
and (2=N)

N
m=1
( )
sin c
m
. This is performed after the periodic solution
has been obtained, over one period of rotation. The
resulting equations will be in the non-rotating frame
and contain only higher harmonics (3/rev and
higher). By averaging over one period, the nal
constant coefcient system is obtained in the form
of (4.1.3). For the blades y will contain: the
corresponding q
0
; q
c
and q
s
DOF and their time
derivatives for both the purely structural and the
aerodynamic DOF [171]. In the above approach it is
worth noticing that the averaging procedure will
force the reference state to appear in the equations.
This is an important aspect when dealing with non-
linear systems.
Stability of wind turbines has been considered
systematically in Europe in the late 1990s [176], while
today there is an on-going activity under the EU
funded project STABCON. Several linear stability
tools have been produced along the lines described
above and good correlation has been obtained
amongst the different codes, see [177] and the
references cited. Inevitably linear theory is approx-
imate so the results it produces are subject to cross
checking. Clearly utter measurements are difcult
to obtain because of the involved risk, therefore it is
indispensable to rely on theoretical developments
which will be discussed in the next section.
4.2. Aeroelastic coupling: linear vs. non-linear
formulations
Linearization of the aeroelastic equations cer-
tainly offers computational efciency. However, it is
not always appropriate. For the structure, linear
theory requires that the deformations and displace-
ments are small; an assumption not always valid.
For the aerodynamics and its coupling with the
structure, linearization will suppress some depen-
dencies involved in (4.1.1) and therefore inuence
the estimation of the aerodynamic damping. In this
connection the use of semi-empirical unsteady
aerodynamic models introduces still unresolved
uncertainties. Furthermore there are cases in which
the multi-blade transformation is not applicable e.g.
when blades are non-identical, or when the rota-
tional speed varies.
Retaining fully the non-linearities of the aero-
elastic problem has the following consequences: the
structure is considered at its deformed state, the
coupling conditions among the components are
ARTICLE IN PRESS
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 309
taken in their full form, and the aerodynamic
equations are not simplied. The last point is
important with respect to the best possible estima-
tion of the aerodynamic damping within the
margins offered by the semi-empirical models used
in the currently available aeroelastic tools. In this
respect multi-body analysis, which has been
adopted in most of the existing stability tools
[161,178181], offers this option by construction,
provided that the structural model considers the
construction at its deformed state and the code
takes this into account in the denition of the local
aerodynamics.
Under such conditions, the information of the
stability characteristics of the wind turbine are
contained in the time signals of the loads calculated
through a non-linear time domain simulation. There
exist two types of methodologies for retrieving this
information: the one is based on the work-based
approach [176], while the other is using signal-
processing techniques. In the work-based method
any mode shape is excited and loads are recorded
until a steady periodic state has been reached. Then
the aerodynamic damping is estimated by the work
of the aerodynamic loads over the last period under
the assumption that there is no energy interchange
between the modes through the aerodynamic loads.
In the signal processing approach, the system is
harmonically excited for a nite duration. Then the
transient response following the abrupt termination
of the excitation is recorded wherefrom the aero-
elastic damping and frequency of the specic mode
are determined. There are two different methods,
both well documented in the literature [182184]
that can be used: the moving block method and the
method of Hilbert transform. The moving block
method is a FFT based method, commonly used in
rotorcraft applications [182,184]. The transient
response amplitude is computed on a block of data
using a FFT calculation. The block is then moved
forward by a single point in time and the computa-
tion of the transient response amplitude is repeated.
The linear t for the slope of the natural logarithm
of the sequence of response amplitudes in time
(peak plot) provides the damping (Fig. 21). In the
Hilbert method, the transient response y(t) and its
Hilbert transform ~ y(t) are used to dene the
envelope signal A(t) =

y
2
(t) ~ y
2
(t)
_
[183,184].
For a transient response typical of a viscously
damped system a line can be tted to the logarithm
of this envelope. The slope of this line, as in the case
of the moving block method, gives the damping
(Fig. 22). Comparing the two, it has been found that
the Hilbert method has certain advantages over the
moving block method. In particular the Hilbert
damping analysis provides an estimate of the
decaying envelope signal so it can be used in
assessing the non-linear damping characteristics of
a mode. Therefore it is not limited by the assump-
tion of the viscous damping as the moving block
method does. Moreover, experience has shown that
the Hilbert method is more efcient in calculating
the damping of spectrally close modes [163].
4.3. Examples of time simulations and instabilities
In this section the challenges in aeroelastic design of
wind turbine rotors will be addressed and in particular
the phenomenon of aeroelastic instability of wind
turbine rotors will be explored by showing examples
of simulation results on a wind turbine design.
Back from the beginning of development of
modern wind turbine rotors in the late seventies
there has been concern about the problems that
could arise due to aeroelastic instability. In parti-
cular the use of the stall regulation principle was
uncertain as it was foreseen that a apwise
instability (so-called stall utter) would occur when
operating in the stall region due to the negative
slope of the C
L
vs. a curve. However, it turned out
that the apwise instability during operation in stall
was not the most critical problem, but instead the
edgewise instability, resulting in edgewise blade
vibrations. The rst experimental evidence of this
instability was seen in the mid nineties and initiated
considerable research activities in order to explore
the phenomenon and provide practical solutions.
Below, is mainly focused on this instability as this
serves as a good example to illustrate the problem of
aeroelastic instability in more general terms.
In the more recent wind turbine designs the
regulation of the turbines has shifted from stall
regulation to pitch control where the operational
range for the ow over the blade moves to low angle
of attack at high wind and thus away from the stall
region. This has almost removed the instability
associated with stall during operation but it is still
so that the edgewise blade modes are aerodynami-
cally low damped on the pitch regulated turbines.
A major instability problem on the modern
turbines is seen when the rotor is parked or idling
at very low RPMs at wind speeds above stop wind
speed, which typically is around 25 m/s. Again it is
typically edgewise dominated instability problems
ARTICLE IN PRESS
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 310
ARTICLE IN PRESS
3
5
4
0
4
5
5
0
t
i
m
e

(
s
e
c
)
-
3
.
5
E
+
0
6
-
3
E
+
0
6
-
2
.
5
E
+
0
6
-
2
E
+
0
6
-
1
.
5
E
+
0
6
-
1
E
+
0
6
t o w e r t o p t i l t b e n d i n g m o m e n t ( N m )
1
2
3
4
f
r
e
q
u
e
n
c
y

(
H
z
)
0
5
E
+
0
8
1
E
+
0
8
1
.5
E
+
0
8
2
E
+
0
8
2
.
5
E
+
0
8
3
E
+
0
8
F F T o f t h e t o w e r t o p t i l t b e n d i n g m o m e n t ( N m )

t

d
a
t
a
1
9
1
9
.
0
5
1
9
.
1
1
9
.
1
5
1
9
.
2
1
9
.
2
5
1
9
.
3
1
9
.
3
5
1
9
.
4
0
0
.
5
1
1
.
5
2
2
.
5
3
t

(
s
e
c
)
-

n

l o g o f r e s p o n s e a m p l i t u d e
F
i
g
.
2
1
.
S
c
h
e
m
a
t
i
c
d
e
s
c
r
i
p
t
i
o
n
o
f
t
h
e
m
o
v
i
n
g
b
l
o
c
k
d
a
m
p
i
n
g
a
n
a
l
y
s
i
s
.
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 311
related to stall of the blade. A simulation example
will be shown later to illustrate this problem.
Finally, the utter instability will be addressed. It
seems that the increased up scaling of the turbines
has led to rotor and blade designs where the utter
speed is not so far away from normal operational
speed. Also this instability will shortly be illustrated.
4.3.1. Edgewise blade vibration instability
As mentioned above the rst experimental
evidence of this instability was seen in the mid
nineties on stall-regulated rotors with a diameter of
3540 m. An example is presented in Fig. 23 and it is
seen that the amplitude of the edgewise blade root
moment (which at steady conditions vary sinusoidal
with 1p due to the gravity) increases 23 times due
to instability during operation in stall. The experi-
mental evidence of the edgewise instability led to
considerable research on this subject and a major
European project Prediction of Dynamic Loads
and Induced Vibrations in Stall funded by the EU
was carried out in the period from 19951998 [176].
ARTICLE IN PRESS
-1.5e+06
-1e+06
-500000
0
500000
1e+06
1.5e+06
2 4 6 8 10 12 14
t
o
w
e
r

t
o
p

t
i
l
t

b
e
n
d
i
n
g

m
o
m
e
n
t

(
K
N
m
)
time (s)
transient signal
envelope signal
envelope signal exp. fit
Fig. 22. Envelope signal calculation of the tower top tilting moment transient response using the Hilbert transform method for the same
signal as in Fig. 21.
1000
800
600
400
200
0
-200
0 50 100 150 200 250 300
Time [sec.] Time [sec.]
E
D
G
E
W
I
S
E

B
L
A
D
E

R
O
O
T
M
O
M
E
N
T

[
U
N
C
A
L
I
B
R
A
T
E
D
]
E
D
G
E
W
I
S
E

B
L
A
D
E

R
O
O
T
M
O
M
E
N
T

[
U
N
C
A
L
I
B
R
A
T
E
D
]
EDGEWISE BLADE VIBRATION INSTABILITY EDGEWISE BLADE VIBRATION INSTABILITY
--MEASUREMENT
1000
900
800
700
600
500
400
300
200
100
0
-100
200 210 220 230 240 250 260
Fig. 23. Measured edgewise blade root moment on a stall regulated rotor at high wind. Detail of the time track to the right.
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 312
The origin of the instability is simple. If a rotating
airfoil section is harmonically translated along an
axis x
B
and the direction of this axis y
RB
relative to
the orientation x
R
of the rotor plane is varied, the
aerodynamic damping coefcient for the section
varies considerably, see Fig. 24 from [185]. For low
y
RB
which means in-plane or edgewise vibration
direction the damping coefcient is negative even at
low wind speed. For vibrational directions close to
901, which corresponds to out-of-plane or apwise
direction the damping coefcient is strongly depen-
dent on the inow wind speed. It is highly damped
at low wind but close to zero or negative damped at
high wind.
On a complete rotating blade the direction of
vibration depends on the structural design of the
blade as well as of the complete turbine structure.
The movement of the tip section of the blade will
now no longer necessarily be along a straight line.
However, typically the tip section of the blade in the
rst apwise mode will be on a path almost
perpendicular to the rotor plane and in the 1st
edgewise mode it will almost be in-plane. It is thus
expected that the basic damping characteristics of
ARTICLE IN PRESS
Fig. 24. The aerodynamic damping coefcient c_xx_b for a rotating airfoil section as function of vibration direction y
RB
and at three
different inow velocities, from [185].
Fig. 25. Damping characteristics computed for a rotating, isolated blade using the linear aeroelastic stability tool HAWCStab [186].
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 313
the whole blade will be comparable with the
characteristics of a single airfoil section. Using the
linear stability tool HAWCStab the damping
characteristics for a rotating blade can easily be
derived and the computed damping characteristics
for a rotating 19 m blade (500 kW rotor) corre-
sponding to the blades of the rotor where the
experimental instability was shown in Fig. 23, are
shown in Fig. 25 from [186]. It is seen that the
damping for the rst apwise mode varies from
highly positive values at low wind speed and to
values close to zero at 14 m/s and then increasing
slightly at higher wind speeds. It should be noted
that the inuence of using unsteady blade section
aerodynamics is shown (modelled with the Beddoes
Leishman model in the present case) and that this
has considerable inuence on the aeroelastic damp-
ing. Generally, the unsteady aerodynamic effects
increase the damping. The damping for the rst
edgewise mode is seen to decrease gradually with
increasing wind from being slightly positive at low
wind to slightly negative at high wind. Finally, it
should also be noted that the damping shown is the
total damping including the structural damping
which in the present case is around 2% for the rst
apwise mode and 3% for the rst edgewise mode.
The nal step in model complexity is now
obtained by going to a full aeroelastic model of
the turbine comprising the dynamics of the shaft,
the nacelle and the tower. The aeroelastic stability
results using HAWCStab is for the complete wind
turbine, see e.g. Fig. 26 from [186], showing the
aeroelastic damping for the rst 10 mode shapes
(numbered from lowest frequency). It is seen that
two mode shapes are negatively damped at high
wind speed and that one mode shape is close to zero
and slightly negatively damped at the highest wind
speed. For comparison, time simulations with the
aeroelastic code HAWC [187,188] is performed on
exactly the same turbine model at a wind speed of 8
and 16 m/s, respectively, in order to see if the
instability at high wind speed predicted by the linear
stability tool can be conrmed by the HAWC
model, with non-linear aerodynamics. The com-
puted edgewise blade root moment shown in Fig. 27
conrm the instability at high wind and compares
very well with the measured instability shown in
Fig. 23. It is thus demonstrated that the edgewise
instability problem can be predicted in both time
simulations and using linearized stability analysis.
However, as will be demonstrated below, the
edgewise vibration instability is much more complex
than just edgewise vibrations of the individual
blades but comprises the dynamic characteristics
of the whole turbine. This understanding is vital for
the development of design solutions preventing the
instabilities.
In order to analyse the instability in more details
the use of parallel aeroelastic time simulations on a
turbine with a stiff structure is made. In this case the
ARTICLE IN PRESS
Fig. 26. Aeroelastic damping computed with the aeroelastic stability HAWCStab [186]. Three modes have negative damping at high wind
speed.
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 314
loads on the different components will mainly
reect the direct load input from the external
aerodynamic forces and from gravity forces. It
should be mentioned that the simulations are
performed at two average wind speeds of 8 and
16 m/s, respectively, and with a turbulence intensity
of 12%.
A power spectral density analysis of the simulated
edgewise blade root moment shows a distinct peak
at 3.07 Hz which is close to the frequency of the rst
edgewise mode for a single, rotating blade which is
3.19 Hz, see Fig. 28. The response at 16 m/s of the
exible turbine at this frequency is seen to be several
orders of magnitude bigger than the response of the
stiff turbine indicating an instability situation.
Somewhat the same tendency is seen at 8 m/s but
with considerable less difference between the re-
sponse of the stiff and exible turbine indicating low
damping but not an instability.
Analysing further the results of the stability
analysis with HAWCStab as shown in Fig. 26 it
can be derived from the results of the model that the
three mode shapes that are negatively damped or
very low damped at 16 m/s are:
+ mode no. 1 with frequency 0.77 Hz and damping
6.05%,
+ mode no. 3 with frequency 0.93 Hz and damping
4.33%,
+ mode no. 7 with frequency 2.73 Hz and damping
1.59%.
ARTICLE IN PRESS
Fig. 28. Comparison of power spectra of edgewise blade root moment for a exible and a stiff turbine, respectively, at 8 and 16 m/s.
Fig. 27. Time simulation with the aeroelastic code HAWC on a 500 kW stall regulated rotor at two wind speeds, 8 and 16 m/s.
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 315
The numbering in HAWCStab of the modes is
with increasing mode number with increasing
frequency starting with the lowest frequency. Mode
no. 1 and no. 3 are both modes where lateral tower
bending is the main motion together with edgewise
bending of the three blades and torsion of the main
axis. The edgewise movement of the blades are all in
phase and the difference between the two modes is
the direction of the edgewise bending relative to the
tower bending. Comparing again the results of the
linear stability analysis with the time simulation
results, the instability of the lateral tower bending
dominated modes is conrmed by the spectra of the
tower top and tower bottom lateral bending
moments, Figs. 29 and 30. A distinct peak at
0.79 Hz is seen in both tower top and tower bottom
spectra at 16 m/s and this peak cover probably also
a considerable content at the 0.93 Hz, which was the
frequency of mode no. 3.
It can be noticed that the HAWCStab results do
not contain an unstable mode with a frequency
around 3.07 Hz (the edgewise instability) as seen in
the time simulation results. However, mode no. 7 is
in fact the response on the non-rotating turbine
structure from the edgewise vibrations at 3.07 Hz.
The edgewise vibrations in the present case is a so-
called backward whirling, edgewise mode where
there is a phase shift of 1201, between the movement
of the blades. This whirling mode results in shaft
and tower bending but with a frequency shifted 1p
ARTICLE IN PRESS
Fig. 29. Comparison of power spectra of tower top lateral moment for a exible and a stiff turbine, respectively at 8 and 16 m/s.
Fig. 30. Comparison of power spectra of tower bottom lateral moment for a exible and a stiff turbine, respectively, at 8 and 16 m/s.
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 316
(p is the rotational frequency) up and down,
respectively, compared with the frequency in the
rotating system. As 1p in the present case is 0.45 Hz
a peak in the spectra of tower bending at a
frequency of around 3.070.45 Hz equal to 2.62 Hz
is expected, which is also seen in Figs. 28 and 29.
However, in the HAWCStab results a slightly
higher frequency 2.73 Hz was seen for this mode.
The presented example of the edgewise blade
vibration instability has shown that the low aero-
dynamic damping of an airfoil section undergoing a
motion along a path close to the chordwise direction
can cause instability of quite different modes on the
turbine at the same time. The design challenges with
the objective to minimize the inuence of low
aerodynamic damping or instability are thus big as
e.g. changes in tower design could lead to e.g.
edgewise vibrations on the rotor. A number of
different methods to reduce the risk for edgewise
vibrations on stall-regulated turbines have been
investigated [176,186] and comprises e.g. changes of
the stalling characteristics of the airfoils by so-called
stall strips as well as structural design of the
blades to achieve optimal vibrational directions.
4.3.2. Instability problems of parked rotors
The turbines are normally designed to operate up
to a certain maximum wind speed and above this
wind speed the turbines are shut down. In the
stopped conditions the rotors can be completely
parked or they can idle with low rotational speed
depending on the actual design of the control
system. As part of the certication of the turbine
it must be shown that the turbine can withstand the
wind loads at extreme wind speed conditions with
parked rotor and the wind coming from any
direction in the case that the yawing system of the
turbine is not functioning.
As an example, the simulation of a rotor, parked
at a wind speed of 50 m/s and a yaw error of 601, is
shown in Fig. 31. The turbine is the same as above
and the time track of the edgewise blade root
moment indicates low damping as the amplitudes
vary much in time. To the right in Fig. 31 is shown a
power spectrum of the same signal and it shows that
the blade mainly vibrates in two modes; mode 3 at a
frequency of 0.93 Hz which was described above
and the rst edgewise blade frequency at about
3.19 Hz. For the present inow direction the
vibrations are not unstable but it has to be
documented for all inow directions.
4.3.3. Flutter instability
The last example of an instability is utter. This is
a well-known instability from the aircraft industry
but has not yet been a problem on wind turbines
and has probably not been seen on commercial
turbines. However, with the increasing size of the
blades it seems that the utter speed decreases due
to increasing structural exibility of the blades and
not least the torsional frequency decreases. There-
fore, it is a good idea to include a utter speed
calculation in the design verication for e.g. 50 m
blades and above.
Flutter involves two DOF of the blade; torsion
and translation. The utter speed decreases when
the frequency of these two DOF approach each
other. For 50 m blades the frequency of the rst
ARTICLE IN PRESS
Fig. 31. Simulation of a parked rotor at 50 m/s and a yaw error of 601. Time track of the edgewise blade root moment to the left and a
power spectrum of the same signal to the right.
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 317
apwise mode is typically slightly below 1 Hz and
the frequency of the rst torsional mode will
typically be in the range from 5 to 8 Hz. However,
it has been seen that the utter instability can occur
by a coupling of the second apwise mode and the
rst torsional mode of the blade and of course this is
also a reason for a decreasing utter speed on the
bigger blades. Another important design parameter
for the utter instability is the centre of mass in the
blade sections relative to the centre of the elastic
axis. As the centre of mass moves away from the
elastic axis in the direction of the trailing edge the
utter speed decreases.
An example of a simulated utter instability is
shown in Fig. 32. The blade does not represent an
industrial blade but has selected structural para-
meters, which give a rather low utter speed. The
second apwise mode for the blade is slightly above
2 Hz and the rst torsional frequency is around
6 Hz. Further the centre of mass is positioned 10%
of the chord length behind the elastic axis. The
example is mainly intended to show the character-
istics of the utter instability, which is completely
different from the instabilities that have been seen
so far. In the simulation the rotor is free to speed up
as the generator is disconnected. When the utter
speed is reached the instability develops within 12 s
and this is completely different from e.g. the
edgewise blade instability that can build up over a
minute or more. This characteristic means the blade
probably will be damaged immediately if the utter
speed is reached. A further consequence of this fast
development is that the utter can occur at a lower
rotational speed of the rotor if there is a yaw error.
This is demonstrated in the right part of Fig. 31
where there is a yaw error of 501. Here, the
rotational speed at utter instability is 13% lower
than if there is no yaw error and the wind speed is
just 8 m/s. If the for example the yaw error of 501
occurred at around the stop wind speed of 25 m/s
the rotational speed at the utter instability would
decrease even further.
5. Present and future developments of aeroelastic
models
The improvement of existing aeroelastic codes
and the development of new aeroelastic models are
highly inuenced by the design trends of new wind
turbines and trends in the siting of the turbines
because this determines the needs for new capabil-
ities of the models.
5.1. Areas with inuence on the development of
aeroelastic models
5.1.1. Inuence of up-scaling
So far the most important design trend has been
the up-scaling which within the last 10 years has
increased the maximum size of the mass produced
turbines with a factor of 10 from about 500 kW with
a rotor diameter of about 40 m to 5 MW turbines
with a rotor diameter of 120 m. The newest turbines
are all pitch controlled with variable speed and
typically with some type of cyclic pitch for load
alleviation. An accurate modelling of these new
exible turbines has increased the requirements to
simulate complex coupled modes where the inte-
grated exibility of the tower, of the drive train and
of the blades and in interaction with the control, is
ARTICLE IN PRESS
Fig. 32. Example of utter instability on a 50 m test blade where the rotor is free to speed up. To the left normal inow whereas to the
right there is a yaw error of 501.
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 318
important. Also non-linear effects from consider-
able deections of the blades and the tower are
becoming important and the aeroelastic stability
must be predicted with a good accuracy. Some
typical frequencies for a 5 MW turbine could be:
+ First tower bending frequency 0.20.25 Hz
+ First apwise frequency 0.81.0 Hz
+ First edgewise frequency 0.91.1 Hz
+ First blade torsional frequency 5.07.0 Hz
The decrease in the fundamental turbine frequen-
cies arising from the up-scaling has the effect that
excitation from the turbulent inow has increased as
the power spectrum of the turbulence peaks at
around 0.05 Hz.
5.1.2. Siting of the turbines
The major part of all new turbines is placed in
small or bigger groups and in some cases in wind
farms with more than 100 turbines. This means that
wake operation is part of the inow conditions,
which has to be simulated in the derivation of the
total design loads of the turbines. Load alleviation
by cyclic or individual pitch controls is important
for such inow conditions and the aeroelastic
models should be suited for this.
Still the major part of turbines are set up on land
but the offshore part will increase, at least if a 510
years interval is considered. On land, the best sites
have been used in many countries and therefore
more and more complex sites will be used. This
means that rather complex inow situations must be
modelled as for example complex shear and non-
uniform turbulence over the rotor disc.
5.1.3. Future trends in turbine design and siting
It seems that the speed of up-scaling might be
slowed down for some years as the industry wants
to be more focused on turbine reliability. One way
to improve this is to have the individual turbine
models on the market for more years than seen in
the past.
However, the increasing offshore development
will support the up-scaling tendency as the tran-
sport and erection of big turbine components is no
longer a major problem. Further, the higher
foundation costs offshore will also support the up-
scaling trend.
Extending the possible offshore sites to deeper
water, new foundation types, typical with multiple
frames, will be developed and the aeroelastic codes
should be able to handle these structures. A further
step in the offshore development is the oating
turbines, either as single oating turbines or more
turbines on the same, oating frame.
Finally, the offshore market could have the effect
that the two bladed turbine with a teetered down-
wind rotor again will be considered as an alternative
to the three-bladed machines. The main barrier for
the development of two bladed turbines has been
the low frequency noise from the blade passage of
the tower shadow. This will be of less importance
offshore as the noise restrictions are much lower
here.
5.2. Areas of development in present and new codes
Inuenced by the continuously increasing require-
ments to the capabilities of the aeroelastic models as
described above, existing models are being further
developed and improved and new codes are written.
Below some important areas of development will be
discussed.
+ non-linear structural dynamics,
+ calculation of induction and its dynamics,
+ wake operation,
+ derivation of airfoil data for aeroelastic simula-
tions,
+ complex inow,
+ aerodynamics of parked rotors and
+ off shore turbines including oating turbines.
5.2.1. Non-linear structural dynamics
So far almost all aeroelastic codes have contained
the model assumptions of small deformations and
rotations, but the increased exibility of the turbines
have made uncertainty about the validity of this
assumption for the new designs. The inuence of
non-linear effects on the dynamic response of a
turbine was treated in a recent paper [189] and
different effects were considered. For example
characteristics (frequency and damping) of the rst
edgewise mode will typically change as function of
increased apwise deection because the edgewise
bending will couple more and more with the
torsional mode of the blade. Besides increased
inuence to non-linear effects from the exibility
of the turbine components the new foundations
used offshore and in particular oating foundations
can contribute signicantly to big deections and
related non-linear effects.
ARTICLE IN PRESS
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 319
To overcome this uncertainty about the impor-
tance of non-linear effects a few new non-linear
aeroelastic codes have now been developed. Siemens
wind power has developed a non-linear code
BHawC based on co-rotaing elements [190] while
the new non-linear code HAWC2 [191] developed
by Risoe is based on multibody dynamics.
The trend of including non-linear effects will
continue in the future and codes based on modal
expansion techniques will be developed to include
more modes as e.g. the blade torsional mode.
5.2.2. Calculation of induction and its dynamics
The unsteady aerodynamic air loads are directly
dependent on the computed induction at the same
point on the blade through the angle of attack.
Therefore, an accurate prediction of induction is of
crucial importance.
As the rotor disc has become bigger and bigger
relative to the scales in the turbulent inow,
increasing variation in inow conditions over the
rotor disc are seen. Wind shear contributes to this
variation in inow. The inow variations results in
likewise considerable variation in the loading
expressed through the local thrust coefcient CT
local
over the rotor disc and thus also in induction. The
induction is therefore even in normal operation
highly dynamic and the induction model should be
adapted to these conditions. As an example, Fig. 33
(reproduced from [192]) shows the variation of the
local thrust coefcient at the outer part of the blade
on an 80 m diameter turbine in normal operation.
Another source contributing to the load varia-
tions and thus to the dynamic induction is the
eigenmotion of the blades. This part is also
increasing due to up-scaling and due to the more
exible designs. As an example a blade with a
apwise frequency of 1 Hz as mentioned above and
vibrating with an amplitude of 1 m will experience a
relative velocity component perpendicular to the
chord of around 76 m/s in maximum. This value
can directly be compared with the variations in
inow and is thus considerable. Different develop-
ments to improve the computation of induction
have been seen. One type of modelling is the
coupling of a numerical actuator disc model to the
structural part of the aeroelastic model. So far such
models have mainly been used to verify the accuracy
of BEM modelling. As an example the aeroelastic
code HAWC at Risoe with a BEM-type induction
modelling in the standard version, has been devel-
oped in another version HAWC-3D where the
induction is computed with a 3D actuator disc
model and used e.g. to investigate the accuracy of
yaw modelling [142,193], see Fig. 34 from [142]. In
the coming years aeroelastic codes coupled to an
actuator disc or actuator line model for computa-
tion of induction will certainly be further developed
ARTICLE IN PRESS
Fig. 33. Example of dynamic loading at a point close to the tip of a blade on an 80 m diameter turbine in turbulent wind, from [192].
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 320
but probably mainly used for research and for
verication of simpler codes and not so much in the
aeroelastic codes used by the industry due to the
much longer simulation time.
However, besides a direct coupling of an actuator
disc or line method to an aeroelastic code in order to
compute the induction at each time step there are
other ways to utilize the capabilities of an actuator
disc model. One example is to compute the
induction characteristics of a rotor within its wind
speed operational interval with an actuator disc and
then afterwards use these quasi-steady induction
characteristics in aeroelastic simulations for the
rotor. The BEM method in the new HAWC2
aeroelastic code developed at Risoe has been
implemented in such a way that this method can
be used. The method is shortly rst to compute the
local induction as function of local loading for the
actual rotor in a number of calculation points (n
m) over the rotor with an actuator disc:
a(r
n
; y
m
) = f CT r
n
; y
m
( ) ( ).
Afterwards, these induction functions are then used
in the BEM method during time simulations on the
rotor instead of the unique, single relationship from
1D momentum theory, C
T
= 4a(1 a).
In fact, one of the main deciencies of the BEM
method is that this induction formula concerns the
whole rotor disc. Deviations are seen in regions with
considerable radial variation in the loading, which
means in the tip or root region. Another example
where the momentum induction formula does not
hold is if the rotor disc is not plane, e.g. coned
rotors or rotors where the blades bend considerably.
5.2.3. Wake operation
Most turbines are now set up in clusters or wind
farms and this means that such turbines in part of
their lifetime operate in wake from one or more
turbines. In the international standard for design of
turbines IEC 61400-1 [2] the increased loading from
operation in wake can be taken into account by
using an increased, effective turbulence which
depends on a number of parameters in the
considered wind farm such as e.g. wind turbine
spacing. However, it has turned out that in cases
where more detailed knowledge of the increased
loading on the different turbine components from
wake operation is needed, another more detailed
aeroelastic modelling is needed than just increasing
the turbulence. One such aeroelastic modelling has
been developed over the past few years [194,195].
The main components in the modelling is: (1)
ARTICLE IN PRESS
Fig. 34. Inuence of yaw on the local inow angle. The curves show the difference in local inow angle at yawed operation at 451
compared with non-yawed operation. The HAWC model has a BEM induction model, HAWC-3D a 3D actuator disc model and
EllipSys3D is a full 3D NavierStokes simulation, from [142].
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 321
computation of the wake decit from the up-stream
turbine with an actuator disc model; (2) meandering
of this decit from the big scales in the turbulence
and (3) additional turbulence within the wake. One
of the advantages of this new modelling compared
to the method using an effective turbulence is that
both the mean yaw loads as well as the yaw
dynamics compare well with measurement. Fig. 35
from [194] shows the loads as function of wind
direction and full wake operation at a direction of
around 2101. The highest mean yaw loads occur in
half wake operation at around 1951 and 2201.
Recently, the wake resulting from the interaction of
several turbines in a row was computed in [196] by
combining large eddy simulation of the NS equa-
tions with the actuator line methodology.
It is expected that the more detailed aeroelastic
modelling of wake operation will be developed
considerably in the coming years because such
modelling is necessary for the development of
advanced control algorithms adapted for load
reduction in wakes. The development of the new
detailed models can be in the direction like the
model described above but also using actuator line
models or full 3D rotor models to compute the wake
characteristics, which can be fed into an aeroelastic
simulation.
5.2.4. Derivation of airfoil data for aeroelastic
simulations
The airfoil data used in aerodynamic and aero-
elastic simulations are of crucial importance for the
ARTICLE IN PRESS
Fig. 35. Measured and simulated loads at 8 m/s (crosses and full lines) and at 10 m/s (squares and dashed lines) on a 80 m rotor during
wake operation (full wake at around 2101 on x-axis) using a new aeroelastic wake simulation method taking into account the wake decit
meandering, from [194].
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 322
accuracy of such simulations. This was clearly seen
in the result of the NREL blind test [137] where the
difference in the airfoil data used in the individual
simulation models probably were the most signi-
cant factor for the big variation between the output
of the different simulations. However, it should be
noted that this result is not a representative
indicator of the uncertainty in the industry on
aerodynamic and aeroelastic results on new turbines
as considerable experience has been learned from
older rotor designs in order to adapt airfoil data sets
so that computed rotor power and loads compare
well with measurements. The drawback is that this
introduces conservatism in the design and for
example can hinder the use of new airfoils. There-
fore, methods and models to derive or correct airfoil
data to be used in aeroelastic simulations have been
a key issue for many years in the wind energy
research community and will also be it in the near
future. 2D wind tunnel airfoil data has so far been
the common starting point to set up an airfoil data
set for simulations. In order to correct for so-called
3D ow effects and rotational effects some kind of
correction is typically applied [1921,197]. The
corrections reect the increased lift and drag that
was measured on different rotors in eld rotor
measurements performed 1015 years ago [198,199].
Full 3D CFD rotor computations later conrmed
clearly the tendencies of the 3D ow effects and
rotational effetcts in the form of increased lift in the
blade root region but also increased drag [137139].
With the possibility of running full 3D simula-
tions on a rotor a new source for derivation of 3D
airfoil data sets for input in aerodynamic and
aeroelastic codes has emerged. Methods to extract
the airfoil data from 3D rotor computations have
been developed [200,201]. One example of such a
derived airfoil data set is shown in Fig. 36 from
[202], where a considerable increased maximum lift
on the inboard stations are seen. Within the next
few years it is expected that a considerable effort
will be on further development of the methods to
extract airfoil data from 2D wind tunnel data as
well as 3D CFD rotor data. So far the CFD
computations have mainly been used to extract
quasi-steady data but they will in the future also be
used to tune the parameters in the dynamic stall
models.
5.2.5. Complex inow
Wind turbines are often set up in so-called
complex terrain, which means some kind of
mountainous terrain because the wind potential is
good at such places. However, the consequence is
that the inow conditions can vary considerably
from turbine to turbine due to local variations of
the terrain. As a result extreme wind shear has been
seen to occur over the rotor disc of turbines placed
in such terrain and likewise the turbulence intensity
can be extremely high.
The fast development of CFD codes has now
made it possible to simulate the ow in details
over the terrain spanned by a wind farm in such
complex terrain. The ow data in the form of wind
shear and turbulence can be used as input in
aeroelastic simulations. There is thus a basis for a
simulation tool to micro site turbines not only with
respect to energy production but also loads is
ARTICLE IN PRESS
2
1.8
1.6
1.4
1.2 C
1
C
d
1
0.8
0.6
0.4
0 5 10 15 20 25 30
r=14%
r=28%
r=42%
r=56%
r=69%
r=89%
r=94%
r=97%
r=14%
r=28%
r=42%
r=56%
r=69%
r=89%
r=94%
r=97%
35 40

0.5
0.45
0.4
0.35
0.3
0.25
0.2
0.15
0.1
0.05
0
0
5 10 15 20 25 30 35 40
Fig. 36. Extracted 3D airfoil data in the form of C
L
, C
D
data as function of angle of attack at different radial stations, from a 3D CFD
rotor computation. Figure from [202].
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 323
expected. Thus type of modelling is certainly of
high interest for the wind turbine industry
because considerable costs to repair turbines due
to extreme inow conditions have been experienced
in the past.
5.2.6. Aerodynamics of parked rotors
It is a common procedure to shut the turbines
down at high wind, typically at wind speeds in the
range from 20 to 25 m/s. Ultimate loads on parts of
the wind turbine can occur during standstill condi-
tions at very high wind speed whereas ultimate loads
on other components will occur during operation
below the stop wind speed.
The aerolelastic codes are therefore also used to
compute the turbine response during standstill
conditions and it is particularly the computation
of the aerodynamic loads on the blades that is
uncertain. This is due to the requirement that the
loads on the parked rotor shall be investigated with
the wind from all directions. The models must
therefore be able to compute aerodynamic loads for
angle of attacks in the complete range from 180
o
to
180
o
. Furthermore the inow is turbulent so that it
is highly unsteady aerodynamic loads and a severe
complication is that some inow conditions often
lead to instability with the blade vibrating in a mode
with negative aerodynamic damping. In a recent EU
funded project KNOWBLADE different CFD
codes were used to explore some basic character-
istics of the blade standstill aerodynamics [115].
Because of the complexity of the aerodynamics it is
also believed that the use of CFD will be the main
path to explore the standstill aerodynamics in the
coming years.
5.2.7. Offshore turbines including oating turbines
Presently, there are considerable research and
development efforts on adapting aeroelastic models
to simulate offshore wind turbines, mounted on a
sub-structure standing on the sea-bed, see Fig. 37
from [203], or on a oating foundation.
Different approaches are seen in this develop-
ment. One line is characterized by a complete
integration of the standard aeroelastic model of
the turbine with a hydroelastic model of the sub-
merged supporting structure including the hydro-
dynamic loads, the wave loading and the soil forces
on the part of the support structure in the sea bed
[190,191]. In both these models the complete
structure is described with nite elements and with
engineering sub-models for computation of the
hydrodynamic loads.
Another line of modelling is to couple a
standard aeroelastic model to a separate module
simulating the supporting structure, including wave
loads and hydro loads. This approach has been
presented by Repower [204] where the Flex5
aeroelatic code is coupled with a standard package
ASAS for computations of wave loads on off shore
structures. In the third development line the main
idea is also to use a standard aeroelastic code and
then couple it to a super element of the foundation.
This approach has been presented by Vestas [203]
where the foundation is modelled with a nite
element model but then decomposed to a super
element using the CraigBampton sub-structuring
scheme. Considerable efforts on further develop-
ments of the aeroelastic models for off shore
applications will be seen within the next years and
with more emphasis on deeper water installations.
ARTICLE IN PRESS
Fig. 37. Different support structure concepts from [203].
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 324
6. Discussion
On a wind turbine there is a strong coupling
between the aerodynamic loads and the time-
dependent structural behaviour of the construction.
Statically a blade might change its twist and thus the
angle of attack when deected. But also the angles
of attack are changed when the blades have a
velocity relative to the xed ground. For instance if
the tower is moving upstream and everything else is
stiff it will be felt by the blades as an increased wind
speed and thus higher angles of attack will be
present along the blades. The aerodynamic response
will depend on how the lift and drag vary with the
angle of attack. In stall the lift will decrease yielding
a possible instability, whereas for attached ow the
lift will increase and thus creating a higher load seen
by the blade opposite the movement indicating
stability. On a real wind turbine not only the tower
is vibrating but all other components as well, which
directly feeds back to the angle of attacks and thus
the loads that again alter the motion. To simulate
the aeroelastic response of a wind turbine, a non-
steady structural model including the inertia must
be made. Two methods, i.e. the method of virtual
work applied on modal shape functions and the
FEM, are addressed. To evaluate the aerodynamic
loads, the BEM is still the most widely used due to
its simplicity and computing efciency. To obtain
realistic results some engineering adds-on are
necessary such as the Dynamic Wake, Dynamic
Stall and a yaw model, which are therefore
thoroughly described in this paper. A BEM method
relies, however, on airfoil data and the results are
therefore no better than the input. To avoid the
uncertainties of the engineering adds-on the Actua-
tor Line model can be used, since this method
resolves the physics of these models through the NS
equations. However, the actual blades are not
resolved and to estimate the aerodynamic lift and
drag airfoil data are still needed. To avoid airfoil
data one needs to solve the NS equations and
resolve the blades and the possible boundary layers.
This is extremely computationally costly and there-
fore this model is not likely to replace the BEM
method in the near future. However, the method
can be used to extract airfoil data that can be used
in the less advanced models. It is expected that the
Actuator Line model very soon will replace the
BEM method, since it contains less empirics. With
the AL model the wake is also a part of the solution
and therefore the effect from this wake on a wind
turbine placed further downstream can be calcu-
lated. These simulations are becoming very impor-
tant as wind turbines are grouped in large (offshore)
wind parks. The inviscid ow models are included in
this paper mainly for historical reasons, since they
have played an important role in determining basic
physical features. This type of model was e.g. used
to calibrate the time constants in the Dynamic
Wake model. Also a few Potential ow models exist
that model the induced velocities and the wake
behind a wind turbine by shedding vortex blobs
from the blade surface. This model is similar to the
AL model and may also in some aeroelastic codes
replace the BEM method.
References
[1] Last og sikkerhed for vindmollekonstruktioner, DANSK
STANDARD DS 472 [in Danish].
[2] IEC 61400-1 Ed.3 CD. 2. revision. Wind turbines, Part 1:
design requirements. Edited by IEC TC88-MT1, 2526
May 2004.
[3] Hansen KS. Performance Measurements on two Danish
630 kW WECS. In: Proceedings of the WIND POWER85,
SERI/CP-217-2902, 1985, p. 1305.
[4] Hansen KS, et al. An evaluation of measured and predicted
fatigue loads for the Tjreborg wind turbine. In: Proceed-
ings of ECWEC93 Travemu nde, Germany, 1993,
p. 57982.
[5] Ahlstro m A. Aeroelastic simulation of wind turbine
dynamics. Doctoral thesis in Structural Mechanics, KTH,
Sweden, 2005.
[6] Glauert H. Airplane propellers. In: Durand WF, editor.
Aerodynamic theory. New York: Dover Publications; 1963.
[7] Rasmussen F, Hansen MH, Thomsen K, Larsen TJ,
Bertagnolio F, Johansen J, et al. Present status of
aeroelasticity of wind turbines. Wind Energy 2003;6:21328.
[8] Quarton DC. The evolution of wind turbine design
analysisa twenty year progress review. Wind Energy
1998;1:524.
[9] Hansen AC, Buttereld CP. Aerodynamics of horizontal-
axis wind turbines. Annu Rev Fluid Mech 1993;25:11549.
[10] Shen WZ, et al. Tip loss corrections for wind turbine
computations. Wind Energy 2005;8(4):45775.
[11] Snel H, Schepers JG. Joint Investigation of dynamic inow
effects and implementation of an engineering method.
ECN-C94-107, 1995.
[12] Schepers JG, Snel H. Dynamic inow: yawed conditions
and partial span pitch control. ECN-C-95-056, 1995.
[13] Bramwell ARS. Helicopter dynamics. Paris: Edward
Arnold (Publishers) Ltd; 1976.
[14] Theodorsen T. General theory of aerodynamic instability
and the mechanism of utter. NACA report 496, 1935,
p. 41333.
[15] ye S. Dynamic stall, simulated as a time lag of separation.
In: K.F. McAnulty, editor. Proceedings of the fourth IEA
symposium on the aerodynamics of wind turbines, ETSU-
N-118, 1991.
ARTICLE IN PRESS
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 325
[16] Leishman JG, Beddoes TS. A semi-empirical model for
dynamic stall. J Am Helicopter Soc 1989;34(3):317.
[17] Hansen MH, Gaunaa M, Madsen HA. A BeddoesLeish-
man type dynamic stall model in state-space and indicial
formulations. Risoe-R-1354(EN), 2004.
[18] Johansen J, Srensen NN. Aerofoil characteristics from 3D
CFD rotor computations. Wind Energy 2004;7(4):28394.
[19] Snel H, Houwink B, Bosschers J, Piers WJ, van Bussel
GJW, Bruining A. Sectional prediction of 3-D effects for
stalled ow on rotating blades and comparison with
measurements. In: Proceedings of the ECWEC 1993,
Travemunde, 1993, p. 3959.
[20] Chaviaropoulos PK, Hansen MOL. Investigating three-
dimensional and rotational effects on wind turbine blades
by means of a quasi-3D NavierStokes solver. J Fluids Eng
2000;122:3306.
[21] Bak C, Johansen J. Three-dimensional corrections of airfoil
characteristics for wind turbines based on pressure dis-
tributions. In: Proceedings of the EWEC conference,
Athens, 27 February2 March 2006.
[22] Fuglsang P, Bak C. Status of the Risoe wind turbine
airfoils. In: Proceedings of the EWEC 2003 (CD-ROM
available from EWEA), Madrid, 2003.
[23] Shinozuka M, Jan C-M. Digital simulation of random
process and its applications. J Sound Vib 1972;25:11128.
[24] Veers P. Three-dimensional wind simulation. SAND88-
0152, UC261, Sandia National Laboratories, NM, 1988.
[25] Mann J. Wind eld simulation. Probl Eng Mech
1998;13:26982.
[26] Srensen JN. Three-level viscous-inviscid interaction tech-
nique for the prediction of separated ow past rotating
wing. PhD thesis, AFM-83-03, Technical University of
Denmark, 1986.
[27] Karagiannis F, Simandirakis G, Chaviaropoulos P, Papai-
liou KD. A 3D shear layer prediction method for steady
viscous ows around HAWTs. In: Proceedings of the
ECWEC 1993, Lubeck-Travemunde, Germany, p. 3914.
[28] Milne-Thomson LM. Theoretical aerodynamics. New
York: Dover Publications; 1966.
[29] Richardson SM, Cornish ARH. Solution of three dimen-
sional incompressible ow problems. J Fluid Mech
1977;82:30919.
[30] Joukowski NE. Vortex theory of a rowing screw. Trudy
Otdeleniya Fizicheskikh Nauk Obshchestva Lubitelei
Estestvoznaniya 1912;16:1.
[31] Margoulis W. Propeller theory of Professor Joukowski and
his pupils. NACA Technical Memorandum No. 79, 1922.
[32] Miller RH. The aerodynamic and dynamic analysis of
horizontal axis wind turbines. J Wind Eng Ind Aerodyn
1983;15:32940.
[33] ye S. A simple vortex model. In: Proceedings of the third
IEA symposium on the aerodynamics of wind turbines,
ETSU, Harwell, 1990, p. 4.15.15.
[34] Koh SG, Wood DH. Formulation of a vortex wake model
for horizontal-axis wind turbines. Wind Eng 1991;15(4):
196210.
[35] Wood DH. On wake modelling at high tip speed ratios.
Wind Eng 1992;16(5):291303.
[36] Hess JL. Review of integral equation techniques for solving
potential ow problems with emphasis on the surface
source method. Comput Methods Appl Mech Eng 1975;5:
14596.
[37] Katz J, Plotkin A. Low-speed aerodynamics. New York:
McGraw-Hill; 1991.
[38] Prandtl L, Tietjens OG. Applied hydro-and aeromechanics.
New York: Dover; 1934.
[39] Landhal MT, Stark VJE. Numerical lifting surface
theoryproblems and progress. AIAA J 1977;6(11):
204960.
[40] Kerwin JE, Lee CS. Prediction of steady and unsteady
marine propeller performance by numerical lifting surface
theory. Trans SNAME 1978; 86.
[41] Hess JL. Calculation of potential ow about arbitrary
three-dimensional lifting bodies. McDonnell Douglas
report, MDC J5679-01, 1972.
[42] Rehbach C. Calcul decoulements autour dailes sans
epaisseur avec nappes tourbillonnaires evolutives. Re-
cherche Aerospatiale 1973;2:5361.
[43] Cottet G-H, Koumoutsakos PD. Vortex methods: theory
and practice. Cambridge: Cambridge University Press;
2000.
[44] Bagai A, Leishman JG. Rotor free-wake modelling using a
pseudoimplicit relaxation algorithm. J Aircraft 1995;32(6):
127685.
[45] Bhagwat M, Leishman JG. Accuracy of straight-line
segmentation applied to curvilinear vortex laments. J
Am Helicopter Soc 2001;46(2):1669.
[46] Gould J, Fiddes SP. Computational methods for the
performance prediction of HAWTs. In: Hulle FV,
Smulders P, Dragt J, editors. Wind energy: technology
and implementation. Amsterdam: Elsevier Science Publish-
ers; 1991. p. 2933 [EWEC
/
91].
[47] Robison DJ, Coton FN, Galbraith RAM, Vezza M.
Application of a prescribed wake aerodynamic prediction
scheme to horizontal axis wind turbine in axial ow. Wind
Eng 1995;19(1):4151.
[48] Coton FN, Wang T. The prediction of horizontal axis wind
turbine performance in yawed ow using an unsteady
prescribed wake model. Proc Inst Mech Eng, Part AJ
Power Energy 1999;213:3343.
[49] Afjeh AA, Keith TG. A simplied free wake method for
horizontal axis wind turbine performance prediction. Trans
ASME J Fluid Eng 1986;108:3039.
[50] Simoes FJ, Graham JMR. Prediction of loading on a
horizontal axis wind turbine using a free vortex wake
model. In: Proceedings of the BWEA conference, 1991.
[51] Voutsinas SG, Beleiss MA, Rados KG. Investigation of the
yawed operation of wind turbines by means of a vortex
particle method. In: AGARD conference proceedings, vol.
552, 1995, p. 11.111.
[52] Coton FN, Wang T, Galbraith RAM. An examination of
key aerodynamic modelling issues raised by the NREL
blind comparison. Wind Energy 2002;5:199212.
[53] Voutsinas SG. Vortex methods in aeronautics: how
to make things work. Int J Comput Fluid Dynamics,
2006.
[54] Preuss RD, Suciu EO, Morino L. Unsteady potential
aerodynamics of rotors with applications to horizontal axis
windmills. AIAA J 1980;18(4):38593.
[55] Arsuf G. A general formulation for aerodynamic analysis
of wind turbine. In: Proceedings of the second IEA
symposium on the aerodynamics of wind turbines, Depart-
ment of Fluid Mechanics, Technical University of Den-
mark, 1988.
ARTICLE IN PRESS
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 326
[56] Bareiss R, Wagner S. A hybrid wake model for hawt. In:
McAnulty K, editor. Proceedings of the sixth IEA
symposium on the aerodynamics of wind turbines, ETSU,
Harwell, 1993, p. 7.110.
[57] Mughal B, Drela M. A calculation method for the three-
dimensional boundary-layer equations in integral form.
AIAA Paper 93-0786, 1993, Reno NV, January.
[58] Riziotis VA, Voutsinas SG. Dynamic stall on wind turbine
rotors: comparative evaluation study of different models.
In: Proceedings of the EWEC97, 1997, Dublin.
[59] Chaviaropoulos PK, Nikolaou IG, Aggelis KA, Srensen
NN, Johansen J, Hansen MOL, et al. Viscous and
aeroelastic effects on wind turbine blades. The VISCEL
Project, Part I: 3D NavierStokes rotor simulations. Wind
Energy 2003;6(4):36585.
[60] Rankine WJM. On the mechanical Principles of the Action
of Propellers. Transactions of the Institution of Naval
Architects, vol. 6, 1865.
[61] Froude RE. On the part played in propulsion by difference
of uid pressure. Trans R Inst Naval Arch 1889;30:
390405.
[62] Wu TY. Flow through a heavily loaded actuator disc.
Schiffstechnik 1962;9:1348.
[63] Greenberg MD, Powers SR. Nonlinear actuator disc theory
and ow eld calculations, including nonuniform loading.
NASA CR 1672, NASA, 1970.
[64] Greenberg MD. Nonlinear actuator disc theory. Z Flug-
wissensch 1972;20(3):908.
[65] Conway J. Analytical solutions for the actuator disc with
variable radial distribution of load. J Fluid Mech 1995;
297:32755.
[66] Conway J. Exact actuator disc solution for non-uniform
heavy loading and slipstream contraction. J Fluid Mech
1998;365:23567.
[67] Madsen HA. The actuator cylinder ow model for vertical
axis wind turbines. PhD dissertation, Aalborg University
Centre, 1982.
[68] van Kuik GAM. On the limitations of Froudes actuator
disc concept. PhD thesis, Eindhoven University of Tech-
nology, Netherlands, 1991.
[69] Srensen JN, Myken A. Unsteady actuator disc model for
horizontal axis wind turbines. J Wind Eng Ind Aerodyn
1992;39:13949.
[70] Srensen JN, Kock CW. A model for unsteady rotor
aerodynamics. J Wind Eng Ind Aerodyn 1995;58:
25975.
[71] Srensen JN, Mikkelsen R. On the validity of the blade
element momentum theory. In: Helm P, Zervos A, editors.
Proceedings of the 2001 European wind energy conference
and exhibition, WIP-renewable energies, Munchen, 2001,
p. 3626.
[72] Madsen HA. A CFD analysis for the actuator disc ow
compared with momentum theory results. In: Pedersen B,
editor. Proceedings of the 10th IEA symposium on the
aerodynamics of wind turbines, Department of Fluid
Mechanics, The Technical University of Denmark, 1996,
p. 10924.
[73] Srensen JN, Shen WZ, Munduate X. Analysis of wake
states by a full-eld actuator disc model. Wind Energy
1998;1:7388.
[74] Madsen HA, Rasmussen F. The inuence of energy
conversion and induction from large blade deections. In:
Proceedings of the European wind energy conference,
James & James, 1999, p. 13841.
[75] Mikkelsen R, Srensen JN, Shen WZ. Modeling and
analysis of the ow eld around a coned rotor. Wind
Energy 2001;4:12135.
[76] Fejtek I, Roberts L. NavierStokes computation of wing/
rotor interaction for a tilt rotor in hover. AIAA J
1992;30(11):2595603.
[77] Rajagopalan RG, Mathur ST. Three dimensional analysis
of a rotor in forward ight. J Am Helicopter Soc
1993;38(3):99.
[78] Masson C, Smali A, Leclerc C. Aerodynamic analysis of
HAWTs operating in unsteady conditions. Wind Energy
2001;4(1):122.
[79] Masson C, Ammara I, Paraschivoiu I. An aerodynamic
method for the analysis of isolated horizontal-axis wind
turbines. Int J Rotating Mach 1997;3:2132.
[80] Hansen MOL, Srensen NN, Flay RGJ. Effect of placing a
diffuser around a wind turbine. Wind Energy 2000;4(3):
20713.
[81] Phillips D, Schaffarczyk AP. Blade-element and actuator
disc models for a shrouded wind-turbine. In: Thor S-E,
editor. Proceedings of the 15th IEA symposium on
aerodynamics of wind turbines, FOI, Swedish Defence
Research Agency, 2001. p. 95105.
[82] Mikkelsen R, Srensen JN. Modelling of wind tunnel
blockage. In: Thor S-E, editor. Proceedings of the 15th IEA
symposium on aerodynamics of wind turbines, FOI,
Swedish Defence Research Agency, 2001, p. 4151.
[83] Mikkelsen R, Srensen JN. Yaw analysis using a numerical
actuator disc model. In: Pedersen B, editor. Proceedings of
the 14th IEA symposium on the aerodynamics of wind
turbines, Department of Fluid Mechanics, Technical
University of Denmark, 2000, p. 539.
[84] Masson C. Viscous differential/actuator disc method and
its applications. In: Thor S-E, editor. Proceedings of the
15th IEA symposium on aerodynamics of wind turbines,
FOI, Swedish Defence Research Agency, 2001, p. 6580.
[85] Ammara I, Leclerc C, Masson C. A viscous three-
dimensional differential/actuator disc method for aerody-
namic analysis of wind farms. J Solar EnergyTrans
ASME 2002;124(4):34556.
[86] Srensen JN, Shen WZ. Numerical modelling of wind
turbine wakes. J Fluids Eng 2002;124(2):3939.
[87] Leclerc C, Masson C. Towards blade-tip vortex simulation
with an actuator-lifting surface model. AIAA-2004-0667,
2004.
[88] Mikkelsen R. Actuator disc methods applied to wind
turbines. PhD dissertation, Department of Mechanical
Engineering, DTU, Lyngby, 2003.
[89] Ivanell SSA. Numerical computations of wind turbine
wakes. PhD dissertation, KTH, Royal Institute of Tech-
nology, Stokholm, 2005.
[90] Arieli R, Tauber ME. Computation of subsonic and
transonic Flow about lifting rotor blades. AIAA Paper
79-1667, 1979.
[91] Borland C, Rizzettaq D, Yoshihara H. Numerical solution
of three-dimensional unsteady transonic ow over swept
wings. AIAA Paper 80-1369, 1980.
[92] Sankar NL, Malone JB, Tassa Y. An implicit conservative
algorithm for steady and unsteady three-dimensional
potential ows. AIAA Paper 81-1016, 1981.
ARTICLE IN PRESS
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 327
[93] Steger JL, Caradonna FX. Conservative implicit nite
difference algorithm for the unsteady transonic potential
equations. AIAA Paper 80-1368, 1980.
[94] Caradonna FX, Tung C, Desopper A. Finite difference
modeling of rotor ows including wake effects. Journal of
the American Helicopter Society, April, 1984, p. 2633.
[95] Sankar NL, Wake BE, Lekoudis SG. Solution of the
unsteady Euler equations for xed and rotor wind
congurations. J Aircraft 1986;23(4).
[96] Jameson Baker TJ. Solution of the Euler equations for
complex congurations. AIAA Paper 83-1919, 1983.
[97] Pulliam TH. Euler and thin layer NavierStokes codes:
ARC2D, ARC3D, notes for computational uid dynamics
users workshop, The University of Tennessee Space
Institute, March 1216, 1984.
[98] Agarwal RK, Deese JE. Euler calculations for oweld of a
helicopter rotor in hover. J Aircraft 1997;24(4).
[99] Srinivasan GR, McCroskey WJ. NavierStokes calcula-
tions of hovering rotor owelds. J Aircraft 1988;25(10).
[100] Srinivasan GR, Raghavan V, Duque EPN. Floweld
Analysis of modern helicopter rotors in hover by Navier-
Stokes method. International Technical Specialists meeting,
Rotorcraft Acoustics and Rotor Fluid Dynamics, October
1517, 1991, Philadelphia, Pennsylvania, USA.
[101] Srinivasan GR, Baeder JD, Obayashi S, McCroskey WJ.
Floweld of a lifting rotor in hover: a NavierStokes
simulation. AIAA J 1992;30(10).
[102] Berkman ME, Sankar LN, Berezin CR, Torok MS.
NavierStokes/full potential/free-wake method for rotor
ows. J Aircraft 1997;34(5).
[103] Hansen MOL, Srensen JN, Michelsen JA, Srensen NN.
A Global NavierStokes Rotor prediction model. AIAA
97-097, 1997.
[104] Srensen NN, Hansen MOL. Rotor performance predic-
tions using a NavierStokes method. AIAA 98-0025, 1998.
[105] Xu G, Sankar LN. Computational study of horizontal axis
wind turbines. AIAA 99-0042, Reno, NV, January 1999.
[106] Duque EPN, van Dam CP, Hughes S. NavierStokes
Simulations of the NREL combined experiment Phase II
rotor. proceedings 1999, ASME wind energy symposium,
37th AIAA Aerospace Science Meeting and Exhibit, AIAA
99-0037, Reno, NV, January 1999.
[107] Srensen NN, Michelsen JA. Aerodynamic predictions for
the unsteady aerodynamics experiment phase-II rotor at
the National Renewable Energy Laboratory. AIAA-2000-
0037, 2000.
[108] Srensen JN. VISCWIND, Viscous effects on wind turbine
blades. ET-AFM-9902, Department of Energy Engineer-
ing, Technical University of Denmark, June 1999,
ISBN:87-7475-218-9.
[109] Chaviaropoulos PK, Nikolaou IG, Aggelis K, Srensen
NN, Montgomerie B, von Geyr H, et al. Viscous and
aeroelastic effects on wind turbine blades: the Viscel
Project. European wind energy conference, Copenhagen,
Denmark, 26 July, 2001.
[110] Kang S, Hirsch C. Features of the 3D viscous ow around
wind turbine blades based on numerical solutions. Eur-
opean Wind Energy Conference, Copenhagen, Denmark,
26 July, 2001.
[111] Johansen J, Srensen NN, Reck M, Hansen MOL,
Stuermer A, Ramboer J, Hirsch C, Ekaterinaris J,
Voutsinas S, Perivolaris Y. KNOW-BLADE task-3.3
report: Rotor blade computations with 3D vortex gen-
erators. Ris-R-1486(EN), 2005, 65p.
[112] Srensen NN, Johansen J, Conway S, Voutsinas S, Hansen
MOL, Stuermer A. KNOW-BLADE task-3.2 report: tip
shape study. Ris-R-1495(EN), 2005, 49p.
[113] Politis ES, Nikolaou IG, Chaviaropoulos PK, Bertagnolio
F, Srensen NN, Johansen J. KNOW-BLADE Task-4
report: NavierStokes aeroelasticity. Ris-R-1492(EN),
2005, 39p.
[114] Johansen J, Srensen NN, Zahle F, Kang S, Nikolaou I,
Politis ES, et al. KNOW-BLADE Task-2 report: Aero-
dynamic accessories. Ris-R-1482(EN), 2004, 33p.
[115] Srensen NN, Johansen J, Conway S. CFD computations of
wind turbine blade loads during standstill operation KNOW-
BLADE, Task 3.1 report. Ris-R-1465(EN), 2004, 28p.
[116] Buning P, et.al. OVERFLOW user manual ver. 1.6ap.
[117] Meaking R. Moving grid overset grid methods for complete
aircraft tiltrotor simulations. AIAA Paper 93-3350, July
1993.
[118] Rizzi A, Eliasson P, Lindblad I, Hirsch C, Lacor C,
Haeuser J. The engineering of multiblock/multigrid soft-
ware for NavierStokes ows on structured meshes. J
Comput Fluids 1993;22:34167.
[119] Kroll N, Radespiel R, Rossow CC. Structured grid solvers.
I: accurate and efcient ow solvers for 3D applications on
structured meshes. AGARD report 807, 1995.
[120] Eliasson P. Edge, a NavierStokes solver for unstructured
grids. FOI-R-0298-SE, 2001.
[121] Chorin AJ. A numerical method for solving incompressible
viscous ow problems. J Comput Phys 1967;2:1226.
[122] Rogers SE, Kwak D. Steady and unsteady solutions of the
incompressible NavierStokes equations. AIAA J 1991;
29(4).
[123] Harlow FH, Welch JE. Numerical calculations of time-
dependent viscous incompressible ow of uid with free
surface. Phys Fluids 1965;8:2182.
[124] Chorin AJ. Numerical solution of the NavierStokes
equations. Math Comput 1968;22(104):74562.
[125] Patankar SV, Spalding DB. A calculation procedure for
heat, mass and momentum transfer in three-dimensional
parabolic ows. Int J Heat Mass Transfer 1972(15):1787.
[126] Rhie CM. A Numerical study of the ow past an isolated
airfoil with separation. PhD thesis, University of Illinois,
Urbane-Champaign, 1981.
[127] Menter FM. Zonal Two equation ko Turbulence models
for aerodynamic ows. AIAA-paper-932906, 1993.
[128] Spalart PR, Allmaras SR. A one-equation turbulence
model for aerodynamic ows. AIAA-92-0439, 1992.
[129] Baldwin BS, Barth TJ. A one-equation turbulence trans-
port model for high Reynolds number wall-bounded ows.
AIAA 91-0610, 1991.
[130] Baldwin BS, Lomax H. Thin layer approximation and
algebraic model for separated turbulent ow. In: AIAA
16th aerospace science meeting, Huntsvill, Alabama, 1978.
[131] Travin A, Shur M, Strelets M. Detached-eddy simulation
past a circular cylinder ow. Turbulence Combust
1999;63:293313.
[132] Strelets M. detached eddy simulations of massively
separated ows. AIAA-2001-0879, 2001.
[133] Xu G, Sankar LN. effects of transition, turbulence and yaw
on the performance of horizontal axis wind turbines.
AIAA-2000-0048, 2000.
ARTICLE IN PRESS
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 328
[134] Michelsen JA, Srensen NN. Current developments in
NavierStokes modelling of wind turbine rotor ow. In:
European wind energy conference, Copenhagen, Denmark,
26 July 2001.
[135] Ferziger JH, Milovan P. Computational methods for uid
dynamics. Berlin, Heidelberg, New York: Springer; 2002.
ISBN:3-540-59434-5.
[136] Fingersh LJ, Simms D, Hand M, Jager D, Contrell J,
Robinson M, et al. Wind tunnel testing of NRELs
unsteady aerodynamics experiment. AIAA-2001-0035 Pa-
per, 39th aerospace sciences meeting & exhibit, 2001, Reno.
[137] Simms D, Schreck S, Hand M, Fingersh LJ. NREL
unsteady aerodynamics experiment in the NASA-Ames
wind tunnel: a comparison of predictions to measurements.
NREL/TP-500-29494, June 2001.
[138] Srensen NN. Evaluation of 3D effects from 3D CFD
computations. In: IEA Joint action, aerodynamics of wind
turbines, 14th symposium, Boulder, December 2000.
[139] Srensen NN, Michelsen JA, Schreck S. NavierStokes
predictions of the NREL phase VI rotor in the NASA
Ames 80-by-120 wind tunnel. AIAA-2002-0032, 2002.
[140] Xu G, Sankar LN. Application of a viscous ow
methodology to the NREL phase VI ROTOR. AIAA-
2002-0030, 2002.
[141] Srensen NN, Michelsen JA, Schreck S. Application of
CFD to wind turbine aerodynamics. In: Fourth GRACM
congress on computational mechanics, GRACM 2002,
Patra, 2729 June 2002.
[142] Madsen HA, Srensen NN, Schreck S. Yaw aerodynamics
analyzed with three codes in comparison with experiment.
AIAA-2003-0519, 2003.
[143] Tongchitpakdee C, Benjanirat S, Sankar LN. Numerical
simulation of the aeordynamics of horizontal axis wind
turbines under yawed ow conditions. AIAA-2005-0773,
2005.
[144] Benjanirat S, Sankar LN. Recent improvements to a
combined NavierStokes full potential methodology for
modeling horizontal axis wind turbines. AIAA-2004-0830,
2004.
[145] Duque EPN, Burklund MD, Johnson W. NavierStokes
and comprehensive analysis performance predictions of the
NREL phase VI experiment. AIAA-2003-0355, 2003.
[146] Benjanirat S, Sankar LN. Evaluation of turbulence models
for the prediction of wind turbine aerodynamics. AIAA-
2003-0517, 2003.
[147] Le Pape A, Lecanu J. 3D NavierStokes computations of a
stall regulated wind turbine. In: Proceedings: the science of
making torque from wind, 1921 April 2004, Delft
University of Technology, the Netherlands, 2004.
[148] Johansen J, Srensen NN, Michelsen JA, Schreck S.
Detached-eddy simulation of ow around the NREL
phase-VI rotor. In: Proceedings CD-ROM. CD 2. Eur-
opean wind energy conference and exhibition 2003 (EWEC
2003), Madrid (ES), 1619 June 2003.
[149] Fleig O, Arakawa C. Numerical simulation of wind turbine
tip noise. AIAA-2004-1190, 2004.
[150] ye S. FLEX4 simulation of wind turbine dynamics. In
Proceedings of 28th IEA meeting of experts concerning
State of the Art of Aeroelastic Codes for wind turbine
calculations (Available through IEA), 1996.
[151] Hansen MOL. Aerodynamics of Wind Turbines. James &
James (Science Publishers) Ltd; 2000.
[152] Schepers JG. Verication of European wind turbine design
codes, VEWTDC: nal report. Technical report ECN-C-
01-055, Netherlands Energy Research Foundation ECN,
Petten, 2002.
[153] Chaviaropoulos P. Development of a state-of-the-art
aeroelastic simulator for horizontal axis wind turbines,
Part 1: structural aspects. Wind Eng 1996;20(6):40522.
[154] Hodges DH, Dowell EH. Nonlinear equations of motion
for elastic bending and torsion of twisted non-uniform
blades. NASA report, NASA TN D-7818, 1975.
[155] Hodges DH. A review of composite rotor blade modelling.
AIAA J 1990;28(3):5615.
[156] Kunz DL. Survey and comparison of engineering beam
theories for helicopter rotor blades. J Aircraft 1994;31:
47397.
[157] Jung SN, Nagaraj VT, Chopra I. Assessment of composite
rotor blade modelling techniques. J Am Helicopter Soc
1999;44(3):188205.
[158] Cesnil CES, Hodges DH, Sutyrin VG. Cross sectional
analysis of composite beams including large initial twist
and curvature effects. AIAA J 1996;34(9):191320.
[159] Nim E. Coupling and reduction of the HAWC equations.
RISOE report, RISOE-R-1294 (EN), 2001.
[160] Snel H, Liendenburg C. Aeroelastic rotor system code
for horizontal axis wind turbines: PHATAS-II. In: Pro-
ceedings of the ECWEC 1990, Madrid, Spain, 1990,
p. 28490.
[161] Riziotis VA, Voutsinas SG. GAST: A general aerodynamic
and structural prediction tool for wind turbines. In:
Proceedings of the EWEC97, 1997, Dublin, Ireland.
[162] Hansen AC, Laino DJ. Validation study for AeroDyn and
YawDyn using phase iii combined experiment data. In:
AIAA-1997-943, aerospace sciences meeting and exhibit,
35th, Reno, NV, January 69, 1997.
[163] Riziotis VA, Voutsinas SG. Advanced aeroelastic modeling
of complete wind turbine congurations in view of
assessing stability characteristics. In: Proceedings of the
EWEC06, 2006, Athens, Greece.
[164] Bazoune A, Khulief YA. Shape functions of three-
dimensional Timoshenko beam element. J Sound Vib 2003;
259(2):47380.
[165] Pfeiffer F, Glocker C. Multi-body dynamics with unilateral
contacts. New York: Wiley; 1996.
[166] Bauchau OA. Computational schemes for exible, non-
linear multi-body systems. Multibody System Dynamics
1998;2:169225.
[167] Bauchau OA, Hodges DH. Analysis of non-linear multi-
body systems with elastic couplings. Multibody System
Dynamics 1999;3:16688.
[168] Chaviaropoulos PK, Politis ES, Lekou DJ, Sorensen NN,
Hansen MH, Bulder BH, et al. Enhancing the damping of
wind turbine rotor blades, the Dampblade Project. Wind
Energy 2006;9:16377.
[169] Petot D. Differential equation modelling of dynamic stall.
La Recherche Aerospatiale 1989;5:5972.
[170] Chopra I. Aeroelastic stability of an elastic circulation
control rotor blade in hover. Vertice 1984;8(4):35371.
[171] Chaviaropoulos PK. Flap/leadlag aeroelastic stability of
wind turbine blades. J Wind Energy 2001;4:183200.
[172] Hansen MH. Improved modal dynamics of wind turbines
to avoid stall-induced vibrations. J Wind Energy 2003;6(2):
17995.
ARTICLE IN PRESS
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 329
[173] Friedman PP, Hammond CE, Woo TH. Efcient numerical
treatment of periodic systems with application to stability
problems. Int J Numer Methods Eng 1977;11:111736.
[174] Peters DA. Fast Floquet theory and trim for multi-bladed
rotorcraft. J Am Helicopter Soc 1994;39(4):829.
[175] Coleman RP, Feingold AM. Theory of self-excite mechan-
ical oscillations of helicopter rotors with hinged blades.
NASA report, NASA-TN-3844, 1957.
[176] Petersen JT, Madsen HA, Bjorck A, Enevoldsen P, Oye S,
Ganander H, et al. Prediction of dynamic loads and
induced vibrations in stall. RISO report, RISOE-R-
1045(EN), 1998.
[177] Politis ES, editor. Benchmark calculations on the NM80
wind turbine. CRES, Technical report STABCON project,
2005.
[178] Kirchgassner B. ARLISa program system for analysis of
rotating linear systems. In: Proceedings of the EWEC 1984,
Hamburg, Germany, 1984, p. 226.
[179] Hansen, MH. Stability Analysis of three-bladed turbines
using an eigenvalue approach. In: 42 AIAA aerospace
sciences meeting and exhibit. 2004 ASME wind energy
symposium, Reno, NV, January 2004.
[180] van Engelen, TG, Braam, H. TURBU offshore computer
programme for frequency domain analysis of horizontal
axis offshore wind turbines: implementation. ECN report,
ECN-C04-0479, Petten, The Netherlands, 2004.
[181] van Holten Th, van Overbeek K. Automatic simulation of
complex non-linear dynamic systems. In: 26th European
Rotorcraft Forum, Paper 65, 2000.
[182] Tasker FA, Chopra I. Assessment of transient analysis
techniques for rotor stability. J Am Helicopter Soc
1990;88:3950.
[183] Simon M, Tomlinson GR. Use of the Hilbert transform in
modal analysis of linear and non-linear structures. J Sound
Vib 1984;96:42136.
[184] Smith CB, Wereley NM. Transient analysis for damping
identication in rotating composite beams with integral
damping layers. Smart Mater Struct 1996;5:54050.
[185] Madsen HA, Petersen JT, Bjo rck A, Ganander H,
Winkelaar D, Brand A, et al. Prediction of Dynamic Loads
and Induced Vibrations in StallSTALLVIB. Final
publishable report on Contract JOR3-CT95-0047. Ris
National Laboratory, 1998.
[186] Hansen, MH. HAWCStab aeroelastic stability tool for
wind turbinesusers guide. Ris-I-2232(EN), Ris Na-
tional Laboratory, August 2004.
[187] Petersen, JT. Kinematically nonlinear nite element model
of a horizontal axis wind turbine. PhD thesis, Part 1 and 2.
Ris National Laboratory, Roskilde, Denmark, July 1990.
[188] Petersen, JT. The Aeroelastic Code HawCmodel and
comparison. In: Pedersen BM, editor. Proceedings of state
of the art of aeroelastic codes for wind turbine calculations.
28th meeting of experts, International Energy Agency,
Annex XI. Technical University of Denmark, Lyngby,
April 1112, 1996, p. 12935.
[189] Larsen TJ, Hansen AM, Buhl T. Aeroelastic effects of large
blade deections for wind turbines. In: Proceedings of the
conference: the science of making torque from wind.
Special topic conference Delft 1921 April 2004.
[190] Rubak R, Petersen JT. Monopile as part of aeroelastic
wind turbine simulation code. In: Proceedings of Copenha-
gen Offshore Wind 2005, conference & exhibition, 2628
October 2005.
[191] Larsen JT, Madsen HA, Hansen AM, Thomsen K.
Investigation of stability effects of an offshore wind turbine
using the new aeroelastic code HAWC2. In: Proceedings of
Copenhagen Offshore Wind 2005, 2005.
[192] Madsen HA, Rasmussen F. A near wake model for trailing
vorticity compared with the blade element momentum
theory. Wind Energy 2004;7:32541.
[193] Madsen HA. Yaw simulation using a 3D actuator disc
model coupled to the aeroelastic code HawC. In: Pedersen
BM, editor. Proceedings, 13th symposium IEA joint action
on aerodynamics of wind turbines, Stockholm, 1999.
[194] Thomsen K, Madsen HA. A new simulation method for
turbine in wakesapplied to extreme response during
operation. Wind Energy 2005;8:3547.
[195] Madsen HA, Larsen GC, Thomsen K. Wake ow
characteristics in low ambient turbulence conditions.
In: Proceedings (CD-ROM), Copenhagen offshore wind
conference 2005, Copenhagen (DK), 2528 September
2005.
[196] Troldborg N, Srensen JN, Mikkelsen R. Numerical
simulations of wakes of wind turbines in wind farms. In:
Proceedings of the EWEC 2006, European wind energy
conference, Athens, 2006.
[197] Madsen HA, Rasmussen F. Derivation of three-dimen-
sional airfoil data on the basis of experiments and theory.
In: Proceedings Windpower
/
88, Honolulu, Hawaii, 1988.
[198] Schepers JG, et al. Final report of IEA Annex XIV: Field
rotor aerodynamics. Report ECN-C-97-027, ECN June
1997.
[199] Schepers JG, et al. Final report of IEA Annex XVIII:
enhanced eld rotor aerodynamics database. Report ECN-
C02-016. ECN February 2002.
[200] Hansen MOL, Srensen NN, Srensen JN, Michelsen JA.
Extraction of lift, drag and angle of attack from computed
3-D viscous ow around a rotating blade. In: Proceedings
of the European wind energy conference, EWEC-1997,
Dublin, Ireland, October 1997.
[201] Johansen J, Srensen NN. Method for extracting
airfoil data using 3D CFD computations. In: Thor SE,
editor. IEA Joint Action Committee on Aerodynamics.
Annex IV Aero experts meeting Boulder, CO (US), 56
May 2003, FFA, The Aeronautical Institute of Sweden,
2003.
[202] Johansen J, Srensen NN, Mikkelsen R. Rotoraerodyna-
mik. In the report Ris-R-1434(DA) Forskning i Aero-
elasticitet edited by Christian Bak, February 2004.
[203] Hald T, Hgedal M. Implementation of a nite element
foundation module in Flex5 using CraigBampton sub-
structuring. In: Proceedings of Copenhagen offshore wind
2005, conference & exhibition, 2628 October 2005.
[204] Seidel M, von Mutius M, Ris D, Steudel D. Integrated
analysis of wind and wave loading for complex support
structures of offshore wind turbines. In: Proceedings of
Copenhagen Offshore Wind 2005, conference & exhibition,
2628 October 2005.
ARTICLE IN PRESS
M.O.L. Hansen et al. / Progress in Aerospace Sciences 42 (2006) 285330 330

Das könnte Ihnen auch gefallen