You are on page 1of 24

WIND ENERGY

Wind Energ. 2008; 11:459482


Published online 6 February 2008 in Wiley Interscience
(www.interscience.wiley.com) DOI: 10.1002/we.269

Research A Study on Rotational Effects and


Article Different Stall Delay Models Using
a Prescribed Wake Vortex Scheme
and NREL Phase VI Experiment Data
Simon-Philippe Breton, Department of Civil and Transport Engineering, Norwegian University of
Science and Technology, 7491 Trondheim, Norway
Frank N. Coton, Department of Aerospace Engineering, University of Glasgow, Glasgow G128QQ,
UK
Geir Moe, Department of Civil and Transport Engineering, Norwegian University of Science and Tech-
nology, 7491 Trondheim, Norway

Key words: Understanding of the stall delay phenomenon on wind turbines remains, to this day,
wind turbine incomplete. A correct modelling of this phenomenon, which results from three-dimen-
aerodynamics;
sional rotational effects, is essential in order to make reliable wind turbine simulations
stall delay;
rotational effects; on the basis of two-dimensional airfoil data, such as with the widely used blade element
correction models; momentum method. The present study addresses this issue by testing six existing models
tip effects; intended to correct for stall delay effects, namely those developed by Snel et al., Chavia-
vortex wake ropoulos and Hansen, Raj, Bak et al., Corrigan and Schillings and Lindenburg. For this
simulations
purpose, the models are implemented into a lifting-line-prescribed wake vortex scheme.
Forces along the blades as well as power and root flap bending moment in a head-on
flow configuration are predicted based on these models, and are compared to wind tunnel
data from NRELs phase VI experiment. While load over-prediction in the presence of stall
is in general observed from the use of the different models, significant differences between
the models are still seen. Local over-prediction is generally seen in the tip region, while
discrepancies are obtained, even at low wind speed, for the root flap bending moment.
The results obtained are discussed in terms of deficiencies and strengths of the current
correction schemes, and from there a basis is provided for the development of improved
correction models. Copyright 2008 John Wiley & Sons, Ltd.

Received 15 August 2007; Revised 23 November 2007; Accepted 9 January 2008

Introduction
Accurate predictions of loads are essential for the effective design of wind turbines. An uncertain knowledge
of the loads in fact often leads to systems being over-designed in order to be on the conservative side. This
ultimately leads to an increase in the energy costs. Inadequacies of current simulation methods come in a sig-
nificant proportion from an incorrect modelling of rotational effects affecting the boundary layer on the turbine
blades once the phenomenon of stall is initiated.1,2 These rotational effects are the cause of the so-called stall
delay phenomenon, also refered to as rotational augmentation, which is characterized by significantly increased

* Correspondence to: S.-P. Breton, Department of Civil and Transport Engineering, Norwegian University of Science and Tech-
nology, 7491 Trondheim, Norway. E-mail: simon.breton@ntnu.no

Copyright 2008 John Wiley & Sons, Ltd.


460 S.-P. Breton, F. N. Coton and G. Moe

lift coefficients compared to the corresponding two-dimensional (2-D) case, and by a delay of the occurrence
of flow separation to higher angles of attack.
Ever since this phenomenon was first observed by Himmelskamp on propeller blades,3 it has captured the
attention of many researchers in the helicopter and wind turbine fields, where computational, theoretical and
experimental investigations have been performed. Among others, McCroskey,4 and Savino and Nyland,5 have
performed flow vizualizations on rotating blades, investigating the separated flow along the blade. Milbor-
row6 performed an analytical evaluation of existing experimental data to try and understand the mechanisms
responsible for the stall delay phenomenon. Madsen and Christensen,7 as well as Ronsten,8 performed pressure
measurements on rotating and non-rotating blades in order to quantify the importance of three-dimensional
(3-D) flow effects. Srensen et al.,9 in addition to Narramore and Vermeland,10 used CFD computations to
provide insight into 3-D rotational effects. Dumitrescu and Cardos11 investigated the 3-D effects on the laminar
boundary layer, trying to shed light on the stall delay phenomenon. Very recently, Schreck et al.12 used CFD
computations coupled with experimental data from the NREL phase VI experiment13 to characterize the aero-
dynamic features in the rotating blade boundary layer, as well as to deduce mechanisms responsible for the
rotational augmentation.
The stall delay phenomenon is, however, still far from being completely understood. An illustration of this
comes from the blind comparison exercise14 that followed NRELs phase VI experiment.13 In this exercise,
modellers were asked to predict loads on the wind turbine without having access to the experimental results.
Comparison between experiments and predictions showed very significant discrepancies, even in the case of
steady inflow. Participants concluded that these were most probably because of an incorrect modelling of the
stall delay phenomenon, which is usually done by adjusting 2-D aerodynamic data to account for 3-D rota-
tional effects.
Different models have been developed to correct aerodynamic coefficients for this phenomenon. In a previ-
ous work,15 five such models were investigated using a lifting line, prescribed wake vortex code, in head-on
flow, so as to isolate the stall delay phenomenon. In contrast to a code based on the blade element momentum
(BEM) method, this code models the wake behind the rotor and its induced effect using vortex elements.
Comparison was then made with the very reliable and exhaustive data acquired in the NASA/NREL wind
tunnel.13 Building on conclusions from this previous work, a further study of six existing models, i.e. those of
Snel et al.,16 Chaviaropoulos and Hansen,17 Raj,18 Bak et al.,19 Lindenburg20 and Corrigan and Schillings21 is
reported here. Two new models, i.e. the ones from Raj and Lindenburg, are tested here, while the application
of the models from Snel et al., and Chaviaropoulos and Hansen is modified relative to the previous analysis.
The same simulation code and experimental data are used for the present study. To our knowledge, no such
comprehensive study on these models has been made with the use of a vortex wake code. While stall delay
effects, which are known to be more important near the root of the blade,3 were the focus of this study, tip
effects were also found to be of great concern, and are discussed in detail.

Methods
NREL Phase VI Experiment
In this work, comparison with experimental data measured on a two-bladed 10.1-m-diameter wind turbine
is made. This test series was run in the NASA Ames wind tunnel for NREL, and is referred to as the NREL
unsteady aerodynamics phase VI experiment.13 The turbine used was stall regulated, its blades were twisted
and tapered and the sectional geometry was that of the S809 airfoil. Measurements were performed in a
flow with less than 1% turbulence, and the data obtained are arguably the most reliable and comprehensive
available to this day. Tests were performed in the downwind and upwind configurations, for a wide range of
yaw angles, wind speeds, cone and pitch angles, at a constant rotational speed of 71.6 rpm. This study will
concentrate on the zero yaw angle upwind baseline configuration, in which the loading was almost uniform
for every azimuth angle. Such conditions are optimal to isolate stall delay, and 3-D effects in general. They
are then also ideal to perform tests on different stall delay models. This configuration also has a zero degree

Copyright 2008 John Wiley & Sons, Ltd. Wind Energ 2008; 11:459482
DOI: 10.1002/we
A Study on Rotational Effects and Stall Delay Models 461

Figure 1. Convention on angles, aerodynamic coefficients and forces used. ^: normal to; //: parallel to

cone angle, and a global pitch of 3 degrees, defined as the angle between the chord at the tip of the blade and
the rotational plane. One of the blades was equipped with pressure taps distributed around the airfoil section
at 22 different positions, more concentrated near the leading edge of the section to render a better resolution
in this more active region of pressure distribution. Such a pressure tap distribution was present at five differ-
ent locations on the blade, i.e. at the 30, 47, 63, 80 and 95% spanwise positions. Aerodynamic forces per unit
length at these radial stations could then be found by integrating the measured pressure distribution around
the corresponding airfoil section. In Figure 1, the convention used for the different forces, coefficients and
angles is shown. As no cone angle was used in the studied configuration, all the quantities seen in this figure
are in the same plane. The forces shown will be investigated rather than the corresponding non-dimensional
coefficients, because they tell more about the actual loads acting on the blades. The root flap bending moment
and power were also measured in these experiments. The former was measured using strain gauges located at
the root of the blade, where the flap direction was defined as being normal to the chord at the tip, while the
power was found from the measured low-speed shaft torque multiplied by the rotational velocity. Aerodynamic
estimates for these two quantities were also calculated by NREL from the pressure measurements performed at
the five radial positions. In these calculations, the measured forces per unit length of the blade are considered
constant over an element around each measurement position that extends on both sides halfway to the next
measurement position. The root element extends on one side to 0.25 R, while the tip element extends on one
side to 1.0 R. The forces on each element are found by multiplying the forces per unit length acting on the
element by the length of that element. The root flap bending moment and low-speed shaft torque are obtained
by multiplication of the appropriate force components by the moment arm between the centre of the given
element and the blade root and rotational axis, respectively. The contributions from each element are finally
added together, and the aerodynamic power estimate is found from multiplying the low-speed shaft torque
estimate by the angular velocity of the rotor.

The HAWTDAWG Model


The HAWTDAWG model is a prescribed wake, lifting line code based on vortex theory. In this code, the
blade is modelled as elements which are represented as a line of bound vorticity lying along the blade quarter
chord line, representing the loading on the blade. For wind speeds of 925 m s1, the blade was separated in
16 spanwise elements with a finer distribution towards the tip in order to get a better resolution in this region
of rapidly changing aerodyanamic conditions. For wind speeds of 58 m s1, an optimal distribution of 10

Copyright 2008 John Wiley & Sons, Ltd. Wind Energ 2008; 11:459482
DOI: 10.1002/we
462 S.-P. Breton, F. N. Coton and G. Moe

elements was used which allowed the use of the same vortex core size in the model for every wind speed,
chosen as 1.5% of the blade span. This gives a tip vortex core radius of the order of the airfoil thickness, as
suggested by Leishman.22 If the bound vorticity is changing from one element to the next, trailing vorticity is
introduced to account for the conservation of circulation, in agreement with the theorem of Helmholtz.23 This
trailing vorticity is modelled as series of straight-line vortex filaments progressing downstream of the trailing
edge of the blade. Shed vorticity can also be modelled by this code to account for time-varying bound vorticity.
In the head-on flow case studied here, there is negligible unsteadiness in the incoming flow, giving rise to no
significant contribution to shed vorticity. Shed vorticity can, however, be present in high stalled conditions,
where flow separation leads to vortex shedding into the wake. In such cases, the effect of this shed vorticity
is small in terms of its influence on the global flow field, and can be ignored.
Unlike free wake models, the present method uses prescription functions to specify the wake geometry,
avoiding the calculation of self-induced velocities in the wake. The wake is represented as two distinct regions,
i.e. the near wake and the far wake. The former has the greatest influence on the induced velocities at the
blades, and the latter represents the far field equilibrium condition of the flow. All wake deformation because
of the changes in the velocities in the streamwise and radial directions is confined to the near-wake region.
The prescription of the spatial development of the near wake is achieved by correlation with momentum theory
and consideration of radial induced velocities at the blades. The radial induced velocity variation in the near
wake is expressed as a quadratic function of time. The corresponding axial induced velocity is defined as three
subregions, in each of which the velocity development is approximated as a linear function of time. These
relationships are applied over the entire near-wake region, which is assumed to reach the far-wake state after
a fixed non-dimensional time has elapsed. The corresponding spanwise variation of the far-wake deficiency
function can be expressed, to a good approximation, as a third-order polynomial based on the dimensionless
local radius. These wake prescription functions have been found to be relatively insensitive to realistic changes
in blade planform and solidity and, consequently, the number of blades. The functions and the hybrid free/
prescribed wake scheme used to develop them are described in detail by Robison et al.24
An initial estimate of the induced velocity distribution along the blade is found from BEM theory, from
which the effective velocity, and thereafter the angle of attack distribution, are determined. Airfoil aerodynamic
data consisting of the lift coefficient versus angle of attack are then used to find the lift corresponding to the
computed angles of attack. The KuttaJoukowsky25 law is finally implemented to find the bound vorticity
corresponding to this lift value. The induced velocity distribution, along with the prescription functions, is
employed to develop the wake shape, onto which the trailing vorticity is distributed. Eighteen time steps are
used per revolution, and the wake extended downstream a number of cycles varying from 5 at high wind speed
to 20 at low wind speed, after it was verified for each wind speed that using a higher number of cycles did not
produce a change in the final results. The BiotSavart law26 is thereafter exploited to find the velocities induced
by the wake on the blades. The same scheme as before is then applied, but now using the new induced velocity
distribution from vortex theory. Iterations are performed until global convergence of the wake shape with the
lift distribution is obtained. The angle of attack distribution is finally invoked, along with aerodynamic airfoil
data in the form of lift and drag coefficients as a function of angle of attack, to find the normal and tangential
forces on the blades (see Figure 1). The use of reliable airfoil data is thus seen to be vital in order to obtain
accurate simulation results. The airfoil data employed here were measured in a wind tunnel at Delft University
of Technology (DUT)13 at a Reynolds number of 1 106, which corresponds best to the tests performed in the
NREL wind tunnel, as shown by Coton et al.27 One should also be aware of the fact that the airfoil data might
differ, for the same Reynolds number, from one measurement facility to another, and that the overall results
might depend on the choice of data set used. This is an area which we suggest requires further investigation.
For higher angles of attack than supplied by the available measurements, data measured on a NACA 001228
blade up to 180 degrees were used, relying on the premise that the performance of an aerofoil is fairly shape
independent once it goes into the bluff body state.6
Unlike BEM methods where annular average induction is found, requiring tip and root loss models to
determine local induction at the blade from which the angle of attack can be calculated, the code used here
is based on vortex wake theory and finds induction directly at the lifting line from the effect of the modelled

Copyright 2008 John Wiley & Sons, Ltd. Wind Energ 2008; 11:459482
DOI: 10.1002/we
A Study on Rotational Effects and Stall Delay Models 463

wake. As a result, it is well suited for studies of different stall delay models that intend to correct the airfoil
data for 3-D effects, and tip-and-root effects.
It should be noted that when results obtained using an in-house BEM method code are referred to, this relates
to simulations performed with a simple code based on the theory and method presented in a book by Hansen,29
which implements Prandtls tip loss model, and Glauerts correction for high axial induction.

Existing Correction Models


Six different models to correct the airfoil characteristics for stall delay, including a wide range of different
assumptions, are studied with the use of the present vortex wake lifting line model. The first three models
studied, i.e. those of Snel et al.,16 Chaviaropoulos and Hansen17 and Raj,18 correct the lift and drag coefficients
Cl and Cd for 3-D effects in the following way:
C1,3-D = C1,2-D + gCl Cl
(1)
Cd,3-D = Cd,2-D + gCd Cd
where gCl and gCd are functions inherent in each correction model, and Cl and Cd are the difference between
the the Cl and Cd that would be obtained if the flow did not separate [taken here, respectively, as Cl = 2p*
(a alift=0), and Cd = Cd (a = 0)] and the Cl and Cd measured in a 2-D configuration.
Snel et al.16 proposed a model following from the solution of a simplified form of the 3-D boundary layer
equations on a rotating blade. These simplified equations are found from an order of magnitude analysis of
the boundary layer equations, and are implemented and solved in a computer program, which enforces among
other things transition from laminar to turbulent flow. The method was originally used to compute pressure
distributions and loads for different angles of attack, and at different sections of the blade of a turbine tested
in a 12 16 m wind tunnel in the China Aerodynamic Research and Development Center,8 allowing computa-
tion of 3-D airfoil data for this blade based on the NACA 4418 profile. Comparison with experimental data
suggested accuracy of the code used, and from this a simple correction model was developed by a fit to the
calculated results, which has the form
gCl = 3(c/r )2 (2)
No correction to the drag coefficient was proposed.
Chaviaropoulos and Hansen17 used for their study of the stall delay phenomenon a quasi-3-D model based on
the solution of a system of simplified equations derived by integrating the 3-D incompressible NavierStokes
equations in the radial direction, resulting in mean radial values, which were then subjected to different
assumptions. The model was applied for both laminar and turbulent flows. The equations were numerically
integrated with the use of a pressure correction algorithm, which resulted in pressure distributions around the
airfoil sections from which 3-D values of the lift and drag coefficients were deduced. Comparison with 2-D
coefficients measured in a wind tunnel allowed for the development of a semi-empirical model to correct for
3-D rotational effects. The two most important parameters producing 3-D effects were found to be the ratio
of local chord to local radius and the local twist angle of the blade. These are present in the correction model
proposed, which also uses the idea of Snel et al. of a correction proportional to the factor Cl. In contrast to
the model of Snel et al., a correction for the drag coefficient was also proposed, i.e.

gCl = a(c/r )h cosn ( )


(3)
gCd = a(c/r )h cosn ( )
where f is the local twist angle. The values of the constants given by the authors as a = 2.2, h = 1 and n = 4,
will be used here.
Raj18 proposed a model based on an improvement of the Du and Selig stall delay model.30 The latter was
based on the analysis of the 3-D integral boundary layer equations on a rotating blade, as an extension of the
work of Snel et al.16 As in the Snel et al. model, the boundary layer equations were simplified by an order

Copyright 2008 John Wiley & Sons, Ltd. Wind Energ 2008; 11:459482
DOI: 10.1002/we
464 S.-P. Breton, F. N. Coton and G. Moe

of magnitude analysis. However, the velocity profiles and associated closure relations used to describe the
boundary layer, and necessary for solving the equations, were different. Also, the solution of the equations was
performed differently, as here it was based on an analysis of the effects of rotation on the laminar separation
point. As in Banks and Gadds31 case, an external flow with a linear adverse velocity gradient was assumed;
this quantity representing the rate of change of the external flow velocity going from the leading to the trailing
edge of the blade. Integration of the simplified laminar boundary layer equations up to the point of separation
led to a separation factor representing the position of the separation point on the rotating surface as a function
of two key parameters, i.e. the local chord to radius ratio, and the tip speed ratio. The results obtained for this
factor were approximated by a general equation, which was then used to develop a formula for correction of 2-D
airfoil data to account for rotational effects, based on the same two key parameters, and three additional empiri-
cal correction factors. Raj started from the same correction formula as Du and Selig, and modified it to include
new dependencies, following comparison of results from a BEM method code using his stall delay model with
experimental results from two stall-controlled wind turbines. The most important modification came from the
need for an increase of the drag coefficient because of 3-D rotational effects, as opposed to a decrease which
was predicted by Du and Seligs model. A dependency on the local speed ratio, different from the original tip
speed ratio dependency of the model of Du and Selig, was also put forward, as well as the use of more empiri-
cal constants which were independent of each other for the lift and drag corrections. Raj prescribed values for
these constants based on comparison with experimental power curves. Rajs formulation is also based solely on
Du and Seligs separation factor30 which he says models both the 2-D and 3-D separations, as opposed to the
model of Du and Selig which used the difference between the 2-D and 3-D separation factors. Finally, the new
model includes an added dependency on the radial position, resulting in the following correction formula:
(r/R)nl
al ( c/r )
( )

1 1.6( c/r ) dl r
gCl = 1
2 0.1267 (r/R) nl
R
bl + (c/r )
dl
(4)
(r/R)nd
a ( c /r )
1 1.6( c/r ) dd
( )
d
r
gCd = 2
2 0.1267 (r/R)nd R
bd + (c/r )
dd

Measurement results from the NREL phases III and IV wind turbine tests32 were originally used to determine
the values of the constants associated with this model. Since these tests were performed on a very similar wind
turbine to the one modelled here, we will use the same values for these constants, i.e.
al = 2.0 bl = 0.6 dl = 0.6 nl = 1.0
ad = 1.0 bd = 1.0 dd = 0.3 nd = 1.0
The factors Cl and Cd defined earlier, on which the three presented correction models depend, can become
very high for high angles of attack. It is then important to be careful about how to apply these models in the
eventuality that high angles of attack are reached. In the articles presenting the first two models, no guidance
is provided as to up to which angle of attack the corrections should be applied. Raj18 suggested correcting up
to 30 degrees, after which the application of the correction should be suddenly stopped. In this work, the idea
of Lindenburg20 is implemented such that the models are applied up to a certain angle, chosen here as by Raj
to be 30 degrees, after which the correction itself is linearly decreased to 0, at an angle of attack chosen as
90 degrees, resulting in a smoother transition in the application of the models. An end angle of 30 degrees
was also chosen by Lindenburg33 when applying Snel et al.s stall delay model, and by Laino et al.34 when
applying Du and Seligs model.
Bak et al.19 have recently proposed a model based on the analysis of the difference between the pressure
distribution on a rotating and a non-rotating blade. The phase VI NREL test turbine was used in this process,

Copyright 2008 John Wiley & Sons, Ltd. Wind Energ 2008; 11:459482
DOI: 10.1002/we
A Study on Rotational Effects and Stall Delay Models 465

along with airfoil data measured in a 2-D configuration in a wind tunnel. This difference is expressed via the
quantity Cp representing the contribution from the 3-D effects to the pressure difference between the suction
and pressure sides of the airfoil. This quantity is modelled as the product of a shape function and an ampli-
fication function of the pressure differences. Analysis of the NavierStokes equations for a rotating blade,
coupled with an order of magnitude analysis of the pressure, Coriolis and centrifugal forces, is the basis for
the amplification factor. The shape factor is found from an analysis of the shape of Cp from measurements
at the 30% span location. The resulting model for Cp is given as

( ) ( )()[
2
x f =1
2 2
5 R c
Cp = 1 1+ 1 + tan 2 ( + )] (5)
2 c f = 0 f =1 r r
where x/c is the normalized chordwise position, a is the angle of attack and af=1 and af=0 are, respectively, the
angles of attack where the flow around the airfoil is just about to separate, and just fully separated. Cn and
Ct are then found by integration of this factor along the different blade airfoil geometries, and are added to
the 2-D values to find Cn,3-D and Ct,3-D, the 3-D corrected values of the normal and tangential force coefficients.
The corrected airfoil coefficients Cl,3-D and Cd,3-D are finally found from

Cl,3-D = Cn,3-D cos ( ) + Ct,3-D sin ( )


(6)
Cd,3-D = Cn,3-D sin ( ) Ct,3-D cos( )
It should be noted that the version of Baks model used here is the original one.19 A more recent version of
the model35 implementing a new feature concerning the correction at high angles of attack is now available,
but this was developed in parallel with the work presented here and so is not yet implemented in the current
modelling.
Lindenburg20 proposed a model based on an analysis of the separated flow at the trailing edge. It is described
in terms of the so-called centrifugal pumping mechanism, in which the rotor is acting as a pump on this
separated volume of air, which is then forced to move in the radial direction. The resulting outboard spanwise
flow then undergoes, in a rotating frame of reference, a Coriolis force pushing the flow towards the trail-
ing edge. Separated flow at the trailing edge is associated with the presence of a big suction peak near the
leading edge.36 The amplitude of this peak increases with the square of the radial position, as it is proportional to
the dynamic pressure. The Coriolis loads undergone by the separated flow will act as a favourable pressure gradi-
ent, helping the flow to overcome the negative pressure gradient resulting from the suction peak, and so delaying
separation. An increased normal force on the airfoil is also supposed to follow from the resulting reduction in
volume of the separation bubble. Whereas other authors also consider the centrifugal pumping mechanism in
the development of their model, the latter being for example present in the developed equations via coupled
Coriolis and centrifugal force terms,1618,21 Lindenburgs modelling of this phenomenon is different. Here, the
additional sectional load is expected to be proportional to the size of the separated area of flow. The latter is
supposed to follow the blade angular velocity, and move radially outward with a speed equal to the tangential
velocity r. Considering the Coriolis acceleration affecting the separated area of air, an expression is found for
the additional sectional load, expressed here in terms of a correction to the normal force coefficient:
Cn,3-D Cn,2-D = 1.6(c/r )(1 f )2(cos inf )2 (7)
where c(1 f) expresses the chordwise position of the separation point measured from the trailing edge, and
finf is the inflow angle, equal to the local pitch plus the angle of attack. This correction is directly dependant
on the angle of attack, and is intrinsically reduced for high values of this parameter, for which the cosine term
becomes smaller. The scaling factor 1.6 in this expression was chosen following comparison with the NREL
phase VI torque measurements.20 Note that the larger the separation area is, the more important this correction
becomes. Using the Kirchoff/Helmholtz model,37 the expression for the non-dimensional trailing-edge separa-
tion distance f as a function of the normal force coefficient is given as
Cn = Cn,pot / [(1 + f )/2]2 ( 0 ) (8)

Copyright 2008 John Wiley & Sons, Ltd. Wind Energ 2008; 11:459482
DOI: 10.1002/we
466 S.-P. Breton, F. N. Coton and G. Moe

where a0 is the angle of attack for zero lift, and Cn,pot/a is the slope of the normal force coefficient in the
potential part of the curve. The shift of the separation point towards the trailing edge mentioned earlier, fol-
lowing the centrifugal pumping mechanism, is expressed as a shift in angle of attack. The idea is to assume
the same location for the separation point in the rotating case as for the non-rotating case, but for a higher
angle of attack, given as:
c
rot = 2-D + 0.25rad/(2 ) 1.6
r
(r/Veff )2 () (9)

where Veff is the effective wind velocity seen by the blade, including the tangential and axial induced velocities.
The lift and drag coefficients corrected with this model are found from geometrical components of the normal
force coefficient, using the rotational angle of attack:
Cl,3-D = Cl,2-D + 1.6 (c/r ) (r/Veff )2 [(1 f )2 cos rot + 0.25rad cos ( rot 0 )]

()
c (10)
Cd,3-D = Cd,2-D + 1.6 sin ( rot ) (1 f )2
(r/Veff )2
r
Note that a drag increase is predicted, following the idea that centrifugal pumping of the separated flow requires
energy to be performed. The last term in the expression for C1,3-D concerns scaling which is necessary, follow-
ing the correction of the angle of attack, for the lift coefficient to have the same ratio between the curve for
fully attached and fully separated flows. A particular feature of Lindenburgs model is the special consider-
ation of the flow near the tip of the blade, by applying a different and an independent correction model there,
describing a loss of lift.20 The latter is explained by the idea that the radial flow following from centrifugal
pumping is not driven further than the tip region. This is expected to result in a reduction of underpressure
on the suction side of the airfoil in this region when compared to the non-rotating case, leading to smaller
rotating coefficients. It is noted by Lindenburg that this effect is different than the loss of lift resulting from a
reduction in angle of attack near the tip, which he states can be described in vortex methods by the use of a
tip vortex. This model is to be applied outboard of the 80% spanwise position, and is given by the following
expression for the lift coefficient:
Cl,3-D.tip = Cl,2-D (r/Veff )2 e 1.5( AR )out Cl Cl,2-D /Cl,pot (11)
where (AR)out is the aspect ratio of the part of the blade outboard of the section considered; C1,pot is the potential
lift coefficient, i.e. the one that would be obtained if the flow did not separate; and Cl is the same as defined
earlier. This empirical expression was found using comparisons with blade root flap bending moments and
coefficients measured at the 95% spanwise position on the NREL phase VI turbine. While a reduction of the
drag coefficient was also expected by Lindenburg near the tip because of the same effect of reduction of nega-
tive pressure on the suction side, it has not been modelled yet.
Corrigan and Schillings21 developed their model starting from the analysis of Banks and Gadd,31 based on
the pressure gradients in the boundary layer. Their idea was that the findings of Banks and Gadd could be
used to develop a shape function for stall delay. Banks and Gadds procedure consisted of solving the three
laminar boundary layer equations on a rotating wing, two of them, i.e. the chordwise and radial momentum
equations, being coupled by Coriolis and centrifugal terms. They determined the laminar separation point in
the rotating case by assuming chordwise and radial boundary layer profiles, and imposing an external flow
with a linear adverse velocity gradient K. The angular location on the airfoil surface of the separation point
s, given as the angle between the lines starting at the rotational axis and going, respectively, to the trailing
edge and to the separation point, was found for different values of the non-dimensional velocity gradient K. A
universal relation between this parameter and the angular position of the separation point, in the sense that it
is independent of the blade geometry, was then determined. It was found that laminar separation was delayed
because of the mentioned rotational coupling in the boundary layer equations. Corrigan and Schillings per-
formed the same analysis as Banks and Gadd, and found a 2-D limit value of 0.136 for the term Ks; values
bigger than this limit implying a delay of the laminar separation point towards the trailing edge of the blade.
The universal relation they found was given by

Copyright 2008 John Wiley & Sons, Ltd. Wind Energ 2008; 11:459482
DOI: 10.1002/we
A Study on Rotational Effects and Stall Delay Models 467

s = 0.1517/K 1.084 (12)


The term s was then replaced by the radial chord distribution c/r, which is equivalent to assuming a laminar
separation at the trailing edge of the blade. The product of c/r and the velocity gradient K became a separation
parameter related to the blade geometry, and the following relationship was assumed between this parameter
and the stall delay angle a, which constitutes the correction model:

( )
Kc/r
n
= ( Clmax 0) 1 (13)
0.136
where aC1max is the angle of attack corresponding in this case to the first maximum of the lift coefficient. We see
that the parameter Kc/r is divided by its 2-D limit mentioned earlier. The constant n is related to the intensity
of the rotational effects, and is suggested to be chosen in agreement with test data. Corrigan and Schillings
suggested using a value between 0.8 and 1.6, and mentioned that using n = 1 was satisfactory in many cases.
Tangler and Selig38 also used n = 1 in this model, and this is the value used herein. The whole airfoil table of
non-rotating lift and drag coefficients is to be shifted over this stall delay angle a, where the lift coefficient
is further corrected in the following way:
Cl,3-D( + ) = Cl,2-D( ) + Cl,pot / (14)
where C1,pot/a is the lift slope in the potential region. Note that the drag coefficient finds itself intrinsically
reduced by the shift in angle of attack of the whole airfoil table. Corrigan and Schillings suggested applying
this model up to a 75% of span, above which they inferred that stall delay effects should not be present.

Results
Angle of Attack Distribution and Corrected Airfoil Properties
Figure 2 shows the angle of attack distribution along a blade of the rotating rotor for wind speeds from 5 to
25 m s1, as calculated with HAWTDAWG without the use of a correction model for stall delay. This figure
illustrates the underlying angle of attack distribution on which subsequent calculations of stall delay are based.
It is important to understand that, although stall delay models change the predicted blade loads, the effect that

Figure 2. Angle of attack predicted with HAWTDAWG as a function of position on the blade for different incoming wind
velocities. The wind speed is increasing by 1 m s1 from one curve to the next in the direction indicated by the arrow

Copyright 2008 John Wiley & Sons, Ltd. Wind Energ 2008; 11:459482
DOI: 10.1002/we
468 S.-P. Breton, F. N. Coton and G. Moe

these load changes have on the wake induction and, hence, blade angle of attack distribution is relatively minor.
The only exception to this is in circumstances where very large load changes are predicted in the vicinity of
the root by some of the stall delay models. In such cases, a corresponding local variability in angle of attack
at the root is introduced. It is seen in this figure that angles of attack well above 30 degrees are found for a
wide range of wind speeds, and on a large part of the blade. Figure 3 shows the lift coefficient as a function
of angle of attack based on the 2-D measurements at DUT, and corrected with different stall delay models,
at three different positions on the blade, i.e. at 30, 63 and 95% span. These positions represent the near root,
central and tip areas of the blade. Note that the curves of Rajs correction represent the 25 m s1 case, since
his model depends on the incoming wind speed. Other data, referred to as from NREL measurements, were
derived from the NREL measured values of the tangential and normal force coefficients at five positions of
the blade, and the angle of attack distribution found with HAWTDAWG (Figure 2), using equation (6). A
large spread of the corrected airfoil data is seen. The corrections applied are seen to be larger towards the root,
where 3-D effects are more pronounced.3 Baks model offers the best agreement with NREL measurement
results at the 0.3 R position and above a 25-degree angle of attack. It is difficult to say which model offers

Figure 3. Lift coefficient as a function of angle of attack measured at Delft University of Technology in two-dimensional
tests of blade sections, corrected with different stall delay models, and derived from NREL measurements and the angle
of attack distribution from HAWTDAWG (noted from NREL measurements), for three different positions on the blade,
i.e. (a) 30% span, (b) 63% span and (c) 95% span

Copyright 2008 John Wiley & Sons, Ltd. Wind Energ 2008; 11:459482
DOI: 10.1002/we
A Study on Rotational Effects and Stall Delay Models 469

the best agreement with the NREL data at 0.63 R, because of the large spread in the data. Only the model
from Lindenburg predicts a decrease of the lift coefficient near the tip (Figure 3(c)), as is suggested by the
data derived from the NREL measurements, and at this position his model gives very similar results to the
latter data. As seen in Figure 4, which is the analogue of Figure 3 for the drag coefficient, the data derived
from NRELs measurements suggest an increase of the drag, except for the 95% spanwise position, where
a reduction is predicted over most of the angles of attack. As mentioned before, only the model of Corrigan
and Schillings predicts a decrease of the drag coefficient, except at the 95% spanwise position, where it is not
applied. The models from Bak et al., Chaviaropoulos and Hansen and Raj compare best with the NREL data
at 0.3 R, while Lindenburgs model compares best at 0.63 R. All models compare quite well at 0.95 R, while
none predicts the decrease of Cd suggested by the NREL data at this position.

Power and Root Flap Bending Moment


Figure 5 shows the power measured in the NREL experiment as a function of incoming wind speed, compared
with the prediction from the code HAWTDAWG when using either no correction or different stall delay

Figure 4. Drag coefficient as a function of angle of attack measured at Delft University of Technology in two-
dimensional tests of blade sections, corrected with different stall delay models, and derived from NREL measurements
and the angle of attack distribution from HAWTDAWG (noted from NREL measurements), for three different positions
on the blade, i.e. (a) 30% span, (b) 63% span and (c) 95% span

Copyright 2008 John Wiley & Sons, Ltd. Wind Energ 2008; 11:459482
DOI: 10.1002/we
470 S.-P. Breton, F. N. Coton and G. Moe

Figure 5. Measured power as a function of incoming wind speed compared with power predicted from HAWTDAWG
with and without different stall delay models

models. Good agreement is obtained up to about 10 m s1. For higher wind speeds, it is seen that the non-use
of a model to correct for stall delay results in a considerable under-prediction of the power. However, the use
of most of the correction models leads to a significant over-prediction. Important scatter between the different
models is also observed. Lindenburgs model seems to yield the best agreement overall, but results in under-
predictions at high wind speeds. The effect of the application of Lindenburgs tip reduction of lift model was
checked, and was found to improve the power predictions, which otherwise became too high. Limiting the
application of Corrigan and Schillings model to sections inboard of the 75% span position reduced the over-
prediction from this method. The models of Raj, Chaviaropoulos and Hansen and Snel et al. produced very
significant over-predictions from 10 m s1 and up, while the models from Bak et al., and Corrigan and Schilling
resulted in considerable but smaller over-predictions.
The measured root flap bending moment is compared in Figure 6 with values found by using HAWTDAWG
with and without the stall delay models. It was chosen to investigate this quantity rather than the experimen-
tal thrust, the latter being estimated by integrating force values at the five positions on the blade, and being
expected to be less accurate. Once again, the results from the different models show a wide spread. A remark-
able feature is the systematic difference of about 300 N m between the predictions and measurements at low
wind speeds. That difference was also seen when comparing to results from an in-house simple BEM method
code, which agreed with the HAWTDAWG results. It is to be mentioned that such a discrepancy at low wind
speeds was not observed when comparing the current predictions with the aerodynamic estimate of the root
flap bending moment made by NREL from an integration of the measured forces at five different positions
on the blade. Neither was it observed when comparing the computed thrust with the experimental value from
NREL. The 2-D prediction is seen to be the closest overall to the experimental curve. Lindenburgs and Cor-
rigan and Schillings models lead to good results at high wind speeds, while they result in an over-prediction
at lower wind speeds. Once again, application of Lindenburgs reduction of lift model near the tip was found
to result in a better agreement, its non-use leading to a more significant over-prediction. Not applying Cor-
rigan and Schillings correction model above the 75% position seemed once again helpful. The models of
Raj, and Chaviaropoulos and Hansen lead to the largest over-predictions, especially for wind speeds higher
than 10 m s1, while the models from Bak et al. and Snel et al. also lead to significant but less severe over-
predictions. The power and root flap bending moments are quantities representing integrated forces along the
blade. It is also of great interest to investigate the local distribution of forces. This will be done here using
NRELs pressure measurements performed at five different positions on the blade.

Copyright 2008 John Wiley & Sons, Ltd. Wind Energ 2008; 11:459482
DOI: 10.1002/we
A Study on Rotational Effects and Stall Delay Models 471

Figure 6. Measured root flap bending moment as a function of incoming wind speed compared with root flap bending
moment predicted from HAWTDAWG with and without different stall delay models

Normal and Tangential Forces Along the Blade


Figures 7 and 8 show, respectively, the normal and tangential forces per unit length measured at five positions
along the blade, compared with results from simulations from HAWTDAWG without and with the stall delay
models, for three different incoming wind velocities, i.e. 7, 16 and 25 m s1. The one standard deviation
of the experimental values is indicated around the latter. These three wind speeds cover a range going from
almost no stall to the blade experiencing stall along most of its span, as suggested by Figure 2. At 7 m s1,
the stall delay models have almost no effect on the simulations because of the low angles of attack along the
blade. In Figure 7(a), we see that the normal force is well predicted by the simulations, except near the tip,
where an important discrepancy is seen. The rather small tangential force is quite accurately predicted at this
wind speed, even near the tip, as seen in Figure 8(a). The models of Corrigan and Schillings, and Lindenburg
cause a small over-prediction of the tangential force in the central part of the blade. Disparity between the
results from the different stall delay models appears with an increasing wind speed, along with a worsening of
the agreement with the experimental results. As seen in Figure 7(b), (c), while not using a stall delay model
results in an under-prediction of the normal force at 16 and 25 m s1, except near to the tip, where an over-
prediction is seen, the use of most of the models results in an over-prediction along the blade. The models of
Raj, and Chaviaropoulos and Hansen over-predict the most, but are the only ones that lead to a sufficiently
large force at 16 m s1 near the blade root. At 25 m s1, they also capture the tendency of the force to go up
near the root, while they over-predict its amplitude. Lindenburgs model results in the best prediction along
the blade at 16 m s1, while Bak et al.s, as well as Snel et al.s models, give quite good results at 25 m s1.
However, the former has difficulty, as well as Corrigan and Schillings model, in capturing the rise of the
force near the root, while the latter does not predict strongly enough the decrease of the force near the tip.
Actually, all models except Lindenburgs seem to have difficulty in capturing the behaviour near the tip. In
Figure 7(b), (c) as well as in Figure 8(b), (c), a decrease is seen in Lindenburgs predictions, starting at the
80% spanwise position, and is caused by the transition to the tip model that begins to be applied outboard of
the 80% spanwise position. A similar decrease is seen in these figures for the model of Corrigan and Schillings,
for which the results come in line with the 2-D results from about 75% and outwards, where stall corrections
are no longer applied. As seen in Figure 8(b), (c), an over-prediction of the tangential force generally results
from the use of the stall delay models, whereas the use of no stall delay model led to a prediction not too far
from the experimental result. The over-prediction is more severe towards the root, and is most pronounced

Copyright 2008 John Wiley & Sons, Ltd. Wind Energ 2008; 11:459482
DOI: 10.1002/we
472 S.-P. Breton, F. N. Coton and G. Moe

Figure 7. Normal force per unit length measured at five positions on the blade compared with normal force predicted
from HAWTDAWG with and without different stall delay models, for incoming wind velocities of (a) 7 m s1, (b) 16 m s1
and (c) 25 m s1

in the models of Snel et al., Chaviaropoulos and Hansen and Raj. No stall modelling results in a good agree-
ment at 16 m s1, while the models of Lindenburg and Bak et al. offer the best agreement at 25 m s1. Finally,
it is to be noted that the curve associated with the 2-D predictions, as well as the ones associated with all the
stall delay models except Lindenburgs, shows a tangential force going up near the tip of the blade at a wind
speed of 25 m s1. This tendency was actually observed from a wind speed of 17 m s1, and it was also seen in
simulations obtained from the in-house BEM code, from the same wind speed of 17 m s1. The normal force
simulated with HAWTDAWG and the in-house BEM code was also seen to increase near the tip of the blade,
in the same conditions, from wind speeds of 1724 m s1.

Discussion
The high angles of attack observed in Figure 2 point to the importance of the presence of stall on this blade, and
justify the use of a decaying action when applying stall delay correction models depending on the terms Cl and

Copyright 2008 John Wiley & Sons, Ltd. Wind Energ 2008; 11:459482
DOI: 10.1002/we
A Study on Rotational Effects and Stall Delay Models 473

Figure 8. Tangential force per unit length measured at five positions on the blade compared with normal force
predicted from HAWTDAWG with and without different stall delay models, for wind velocities of (a) 7 m s1,
(b) 16 m s1 and (c) 25 m s1

Cd above, which would otherwise be amplified too much for high angles of attack and lead to too large values
of the aerodynamic coefficients.15 In fact, in Figures 3 and 4, the high angle decay of the correction used for
the models of Raj, Chaviaropoulos and Hansen and Snel et al. particularly at the 30% spanwise position, was
necessary to avoid correcting too much. The spread of the corrected airfoil data seen in these figures follows
from the widely varying assumptions related to each model used. The relatively good agreement at 0.3 R of
the lift and drag coefficients corrected with Baks model with NREL measurement data is consistent with the
use of experimental data at this spanwise position in the empirical modelling of the shape function inherent to
this model. Lindenburgs proposal20 of using a special reduction of the lift force near the tip seems to be quite
useful, as seen in Figure 3(c), as it brings the corresponding lift curve very close to the NREL result. A reduc-
tion of lift near the tip was also predicted by recent theoretical work based on CFD computations.9,39 Corrigan
and Schillings idea of not applying their stall delay correction above the 75% spanwise position also seems
justified, looking at the same figure. There has been some discussion as to how drag is affected by the stall
delay phenomenon (see e.g. Laino et al.34). The drag increase suggested by the NREL measurements in Figure
4, except near the tip of the blade, is in agreement with the most recent theoretical works on the subject.9,39

Copyright 2008 John Wiley & Sons, Ltd. Wind Energ 2008; 11:459482
DOI: 10.1002/we
474 S.-P. Breton, F. N. Coton and G. Moe

Lindenburgs idea33 that centrifugal pumping of air in the separated area requires energy, resulting in a rise of
the drag, seems to be justified. The decrease of the drag coefficient observed near the tip of the blade (Figure
4(c)) could be described by a special model in this region, as suggested for example by Lindenburg,20 but not
implemented. Johansen and Srensen,39 and Srensen et al.9 have also simulated a decreasing drag coefficient
at the 95% spanwise position with the help of a CFD method.
The good agreement seen in Figure 5 between the different power simulations and NREL measured data
at low wind speeds, where stall delay is not of importance, is a sign pointing towards reliability of the code
used. The under-prediction at higher wind speeds when using uncorrected 2-D airfoil data is to be expected
from theory, and shows that 3-D rotational effects should be included in a power simulation. The exaggera-
tion in the simulated power from about 10 m s1 following from the use of certain stall delay models can be
directly related to the over-prediction of the tangential force, at high wind speeds, in the presence of stall,
observed earlier in Figure 8. In fact, the power is calculated by integrating along the blade the product of the
torque force and the distance from the centre of rotation. The torque force is obtained from the projection of
the tangential and normal forces in the plane of rotation (see Figure 1). The contribution from the tangential
force being geometrically more important, the significant over-prediction of the latter following the use of
some stall delay models resulted in too large power predictions at high wind speeds. An overestimation of
the simulated normal force also had an important effect on the power over-prediction. Concerning the relative
effect of the normal and tangential forces on the power, Figure 9 shows the tangential, normal and torque
forces as a function of the position on the blade for a wind speed of 25 m s1, as simulated with HAWTDAWG,
using as an example the model of Chaviaropoulos and Hansen, which resulted in the biggest over-prediction
of the tangential force at this wind speed. The figure also shows the experimental results from NREL. It may
be observed that the over-prediction of the tangential force leads to an important overestimation of the torque
force, particularly near the root, and that results in a significant overestimation of the power. If this force had
been more in line with the experimental result, the torque force would have been very much lower, and the
power, much less overestimated. It is also observed that the contribution of the normal force to the torque
force is not to be neglected. Care should be taken when making comparisons with experimental values of
the tangential force. Tangential force on a blade section is known to be difficult to measure experimentally,
because of the high pressure gradient around the nose of the blade, and the potential for discretely distributed
pressure taps to fail to capture the full extent of the suction peak. Experimental evidence of this difficulty is

Figure 9. Tangential, normal and torque forces per unit length predicted by HAWTDAWG when using the stall delay
correction model of Chaviaropoulos and Hansen, and measured experimentally at NREL, at a wind speed of 25 m s1.
The one standard deviation of the experimental values is indicated around the latter

Copyright 2008 John Wiley & Sons, Ltd. Wind Energ 2008; 11:459482
DOI: 10.1002/we
A Study on Rotational Effects and Stall Delay Models 475

Figure 10. Aerodynamic estimate of the low-speed shaft torque calculated by NREL from force values measured at
five different positions on the blade, compared with low-speed shaft torque measured directly at the shaft. The one
standard deviation of the experimental values is indicated around the latter

obtained by looking at the aerodynamic estimate of the low-speed shaft torque calculated by NREL from the
normal and tangential force values measured at five different positions on the blade. Figure 10 shows this
aerodynamic estimate compared with the value measured directly at the shaft. Above 16 m s1, the estimated
value begins to differ considerably from the value measured at the shaft, and becomes up to 25% less than
the latter. This indicates that the tangential force is indeed difficult to measure on the blade, especially in the
presence of stall, and that the measured values are probably lower than in reality. This suggests then that the
tangential force predictions shown in Figure 8 might not be as far from reality as the figure suggests. In fact,
most of the simulation results shown in the report following NRELs blind comparison exercise14 also exhibited
an over-prediction of the tangential force, becoming more pronounced at higher wind speeds. This illustrates
why comparison of the power with the value stemming from the direct measurement of the low-speed shaft
torque is more reliable.
The outermost elements of the blade have a more important contribution to the power, these elements being
further from the axis of rotation and, therefore, contributing more to the torque moment. As mentioned earlier,
the lift force is expected to go down near the tip of the blade. It is likely that the lift augmentation near the
blade tip, inherent in all the models except Lindenburgs, and Corrigan and Schillings, is in part responsible
for the general over-prediction of power observed. Lindenburgs prediction, which enforces a decrease of the
lift coefficient near the tip of the blade, is the only one that does not result in too high a power. The result-
ing under-prediction with this model at high wind speeds is an indication that the special correction at the tip
may be too strong. Lindenburg mentioned that this model actually needed more refinement.20 The drag coef-
ficient reduction near the tip that he was expecting, but has not yet modelled, would contribute to a rise in the
simulated power, and could possibly push the power curve higher at high wind speeds, bringing it closer to
the experimental one. The observation that the non-application of Corrigan and Schillings model outboard of
75% span resulted in an improvement in the power prediction is also an indication that the lift coefficient on
outboard sections should not be increased by a stall delay correction model. These results show the importance
of an accurate modelling of the behaviour at the tip, and point towards the relevance of using a special airfoil
correction model in this region, where the lift and drag are both expected to decrease.
The significant over-prediction of the power when using Bak et al.s model suggests that it has difficulty
in modelling the effect of stall at larger spanwise positions than 0.3 R, on which its shape function is based.
The over-predictions in power following from the models of Snel et al., Raj and Chaviaropoulos and Hansen

Copyright 2008 John Wiley & Sons, Ltd. Wind Energ 2008; 11:459482
DOI: 10.1002/we
476 S.-P. Breton, F. N. Coton and G. Moe

from 10 m s1 and above, where stall is becoming more important, are indications that these models could
benefit from the decay of the correction beginning at a smaller angle of attack, and that they overestimate the
stall delay phenomenon.
The root flap bending moment is over-predicted in a similar way to the power by the stall delay models.
Once again, the over-prediction of the forces in the tip region contributes more to the moment because of
their greater distance from the blade root. An overestimation of the stall delay effects can obviously also
contribute. However, the current over-prediction is more linked to the exaggeration of the simulated normal
force, which contributes geometrically the most to the root flap bending moment (see Figure 1). In fact, the
models that lead to the biggest over-prediction of the normal force, i.e. those of Raj, Chaviaropoulos and
Hansen and Bak et al., are also the ones that experience the largest over-prediction of the root flap bending
moment. The difference seen at low wind speed for all the predictions is very interesting. It does not depend
on the use of the stall delay models, which do not correct for stall in an appreciable way at these wind speeds.
A similar difference for the root flap bending moment at low wind speeds was observed by other authors (see
e.g. Lindenburg33 and Srensen et al.9). It can also be seen in many of the simulations found in NRELs blind
comparison report.14 This suggests that a systematic error might have been present in the measurements of the
root flap bending moment. This suggestion is reinforced by the fact that the current simulation of the root flap
bending moment agrees well at low wind speed with NRELs aerodynamic estimate despite the over-prediction
of the normal force near the tip of the blade (see Figure 7(a)). Nevertheless, although the over-prediction at
the blade tip does not fully account for the difference in the root flap bending moment, it is inevitable that it
has an adverse effect on the prediction obtained. Figure 11 compares the measured root flap bending moment
with the aerodynamic estimate made by NREL on the basis of the recorded surface pressures. A systematic
difference of about 300 N m is observed between the two curves across the entire wind speed range. If it
were not for that difference, the shape and magnitudes of the two curves would be in close agreement. Unlike
the aerodynamic estimate of the power, the root flap bending moment estimate is mainly based on the normal
force. The calculation of this parameter from measured surface pressures is more reliable than the tangential
force. NRELs aerodynamic estimate shown in Figure 11 is based on five measured force values, considered
constant across each measurement position. Other interpolation techniques could be explored that might lead
to a more reliable estimate, but the recent work of Sant et al.40 may suggest that there is another alternative.
They reconstructed sectional airfoil data from the experiments using a free wake vortex code. On this basis,
they were able to obtain better agreement with the strain gauges measurements.

Figure 11. Root flap bending moment measured with strain gauges at the root of the blade compared with aerodynamic
estimate for this quantity calculated from forces measured at five different positions on the blade. The one standard
deviation of the experimental values is indicated around the latter

Copyright 2008 John Wiley & Sons, Ltd. Wind Energ 2008; 11:459482
DOI: 10.1002/we
A Study on Rotational Effects and Stall Delay Models 477

Assuming there was a systematic error in the measurement of the root flap bending moment, and that the
aerodynamic estimate is more reliable, the simulations conducted here using 2-D sectional aerodynamic data
agree well with the measured data at low wind speed, and underestimate the values at high wind speeds. This
would be in line with expectations when no stall delay effects are considered. The models of Lindenburg,
Corrigan and Schillings and Snel et al. would perform well in comparison to the other models in bringing the
predictions closer to the test data. Lindenburgs model would also, however, lead to a small under-predic-
tion near the tip, suspected to be due once again, in part, to over-correction from his tip model. Corrigan and
Schillings model would also result in a small under-prediction.
Looking in more detail at the local distribution of normal and tangential forces predicted by the different
correction schemes, it appears that no model is capable of consistently predicting the normal and tangential
forces along the blade across the full range of wind speed. An example is the use of Snel et al.s model that
results in reasonable simulations of the magnitude, but not shape distribution, of the normal force along the
blade. This allows acceptable estimates of the root flap bending moment to be obtained but, because the method
significantly over-predicts the tangential force, it significantly overestimates the power. It is expected that the
power prediction could be improved by the introduction of an increased drag coefficient into this model, but
this would have to be done with care.
The quality of the predictions of both the normal and tangential forces at low wind speeds indicates the
reliability of the simulation code used. The slight overestimation of the forces at low wind speeds along the
blade when using the models from Lindenburg, and Corrigan and Schillings, is likely to be caused by too large
an angle of attack correction at these wind speeds. The overestimation of the normal force at 7 m s1 in the
near-tip region is also a feature of the other higher wind speed cases. In fact, most of the predictions presented
in the NREL blind comparison report14 also showed an important overestimation of the normal force in this
region across the wind speed range. The CFD codes, which do not depend on airfoil data input, were the only
ones that did not show such an over-prediction. The fundamental difference between CFD simulations and
the current modelling scheme is that they directly model the local flow topology. They are, therefore, capable
of capturing the full 3-D flow in the vicinity of the tip including induced cross-flow and local vortex suction
effects. This is in addition to the induced downwash from the trailed vortex system that the vortex-based
models do capture.
Shen et al.41 proposed an alternative to existing tip loss correction models to correct airfoil data for local
flow effects near the tip. Their correction was presented in the form of a tip loss model for a BEM code to
replace existing tip loss models. They compared the model to normal force measurements42 at a position as
far out as 0.99 R. The basic idea of Shen et al., even if intended for use in a BEM method simulation code,
is interesting for the present analysis, where some form of correction to 2-D airfoil data near the tip seems
necessary. Johansen and Srensen39 also suggested, based on a CFD analysis, a correction to airfoil data to
account for a decrease of the lift and drag, near the tip of the blade, because of the local effects from the tip
vortex. The decrease was found to exist at even very small angles of attack. However, the behaviour at the
very tip was not shown.
Lindenburgs model implements a special correction in the tip region. However, it is dependant on the angle
of attack, which is small at low wind speeds, and has therefore almost no effect at the lowest wind speed, as
seen in Figure 7(a). Interestingly, however, it does provide some improvement over the other models as the
wind speed increases. The relative success of Lindenburgs model in capturing this effect does suggest that
the use of some form of correction to the airfoil data to account for tip effects is justified. The nature of such a
correction model, however, needs careful consideration, as modelling in this area is known to be very difficult,
and many different criteria have to be fulfilled. As an example, Lindenburgs tip model led to an unrealistic
increase of the normal force near the tip, observed in Figure 7(b). Reasons leading to such an increase, which
was uniquely obtained at 16 m s1, should be analysed carefully in an effort to improve the tip model.
It is important here to differentiate between the corrections for 3-D stall delay on inboard sections of the
blade and those for tip effects. The current predictions suggest that the application of 3-D stall delay correction
methods along the full length of the blade will produce erroneous behaviour in the region of the tip. Similarly,
the results obtained near the tip using Corrigan and Schillings model suggest that simply not correcting the

Copyright 2008 John Wiley & Sons, Ltd. Wind Energ 2008; 11:459482
DOI: 10.1002/we
478 S.-P. Breton, F. N. Coton and G. Moe

airfoil data in this region is not enough to accurately predict the forces there either. Although both stall delay
and tip flow effects are likely to be associated with local 3-D effects, their manifestation is quite different, and
this should be accounted for in any technique that requires corrected 2-D data as input.
The predicted increase of the tangential and normal forces near the tip of the blade is one more indication
that special care needs to be taken in this region. This result can be simply explained. The angle of attack is
going down near the tip of the blade because of the effect of the tip vortex, as seen in Figure 2. If the angle
of attack near the tip is above the 2-D stall angle, then a reduction in the effective incidence will result in an
increase in lift, see e.g. Figure 3(c), which shows the lift coefficient at 0.95 R. Further inboard, where the induc-
tion is less, the effective angle of attack will remain above the stalling incidence and will be associated with
a lower lift coefficient. This result, which we do not think is physically correct, was not obtained when using
Lindenburgs special correction model in the tip region, and can then be avoided by using such a model.
The local force predictions along the length of the blade become worse with rising wind speed, and this
suggests that stall delay effects are not correctly represented by the different models used. None of the models
capture the loading trends along the blade with any level of consistency.
The observed general over-prediction of local loads by the correction models of Chaviaropoulos and Hansen,
Snel et al. and Raj is likely to be due, in part, to the linear dependency of these models on the factor Cl.
Although a decay was imposed on these models above a certain angle of attack in the current implementation,
it may not have been enough. It is likely that the formulation of these first two models could be improved
by the inclusion of a limiting angle above which no correction is made. Raj18 mentioned such an angle in his
formulation, but the present results suggest that this angle may be too high. If this approach is adopted, there
is then an issue over how the data are handled above this maximum angle. One way to add intrinsic decay
at high angles of attack for these models could be to include a local speed-ratio-dependent term, as derived
from the way the centrifugal pumping mechanism was modelled by Lindenburg,33 and used by him also when
applying Snel et al.s correction model.
While modifications of the type discussed earlier may improve the level of prediction, it is unlikely that
they will be universally applicable since the basic modelling strategy may still not properly represent the flow
physics. Looking, for example, at the model of Raj, we see that the basic idea behind the model is that the
forward movement of the separation point with incidence is delayed by 3-D effects. This may well be true,
but measured pressure distributions on inboard sections of wind turbine blades bear little resemblance to 2-D
airfoil pressure distributions with flow separation.43 Indeed, 3-D CFD simulations of such flows indicate that
there are strong spanwise flows set up as separation develops, and these, in turn, influence the chordwise pres-
sure distribution. By failing to represent this aspect of the flow physics, Rajs stall delay model will inevitably
be unable to robustly represent the effects of three-dimensionality across any significant range of flow and
turbine geometries. Corrigan and Schillings model might also suffer problems in its flow modelling for the
same reasons, as it is based on similar premises.
In this study, the model of Raj was chosen over that of Du and Selig,30 from which it was derived, because
it was newer and expected to be an improvement over the latter. However, as Laino and Hansen44 concluded,
it is difficult to declare that Rajs model does better at predicting stall delay than Du and Seligs. In tests
using the same decay of the factors Cl and Cd, Rajs model gave no improvement over Du and Seligs. The
problem might be that Rajs model consists of an empirical fix of Du and Seligs model that did not address
inherent problems identified in Du and Seligs original paper.30 One such problem is that Du and Seligs
model, and also that of Corrigan and Schillings, is based on a laminar boundary layer analysis. Du and Selig
suggested that their model could be improved by incorporating the effect of boundary-layer transition, and
hence, Reynolds number, into the model. Similarly, on the basis of CFD computations, Srensen et al.9 also
suggested that this was necessary.
The fact that many of the existing models were developed using boundary layer theory may be a problem in
itself. For example, Corten36 did a thorough analysis of Snel et al.s model, and concluded that the breakdown
of the boundary layer analysis under separated flow conditions was an inherent problem with the method.
It is likely that fully 3-D viscous computations may provide a more robust basis on which to develop future
correlations.

Copyright 2008 John Wiley & Sons, Ltd. Wind Energ 2008; 11:459482
DOI: 10.1002/we
A Study on Rotational Effects and Stall Delay Models 479

Looking in more detail at some of the features of the existing models, Chaviaropoulos and Hansens
model includes a twist dependency that according to van Rooij and Schepers45 is necessary. The present
results suggest, however, that this feature does not seem to give an improvement over Snels model, which
does not have this dependency but is very similar in other respects. Conversely, the quadratic dependancy
of the correction on the ratio of the chord to spanwise position in Snel et al.s model seems to provide
an advantage over the linear dependancy on the same factor in Chaviaropoulos and Hansens model. It results
in less of an over-prediction of the blade loads because of a reduced correction on outer sections of the
blade.
Bak et al.s model has some interesting features in that it is very empirical and has a dependency on the
angle of attack which, in line with the analysis of van Rooij and Schepers, would appear to be a necessary
feature. As mentioned before, the way this model was developed, with the focus on inboard blade sections,
is probably the cause of its superior accuracy near the root of the blade and deteriorating performance further
outboard. Also, the inexact order of magnitude analysis of the centrifugal and Coriolis forces which provides
the basis for the amplification factor in this model may also be a limitation. The effect from these forces is
central to the modelling of the stall delay phenomenon, and they are known to be difficult to quantify accu-
rately. It is important to note that, in this study, Baks model has been tested against some of the same data
that were used to develop it. In this study, however, the model has been compared to data from across the
entire blade span rather than only with the data at the 30% of span location that was used in its development.
Unfortunately, no other data set including force distributions along the blades was available to test the model
against. It would be interesting to test it against other measurements as they become available, verifying if it
is as general as intended.
Of all the models, Lindenburgs seems to have resulted in the best agreement overall. Its way of model-
ling the centrifugal pumping effect by considering a combination of Coriolis and centrifugal forces acting on
the separated volume of air, and the resulting effect on the chordwise pressure gradient, delay in the occur-
rence of stall and additional normal force, seems to provide some promise. The local speed ratio term in
the model is also very useful at naturally reducing the correction at high angles of attack. The direct depen-
dancy of the model on the angle of attack seems also to be an advantage, as suggested by van Rooij and
Schepers.45
One unique feature of the model of Lindenburg that appears to be particularly beneficial is the special treat-
ment of the tip region. The other models might have given better agreement overall if they had included such a
feature. It is, however, important to point out that they were intended to correct generally for stall delay effects
and did not focus specifically on tip effects. Lindenburgs feature near the tip was developed on the basis of
measurements at 95% of blade span. It may be possible to improve this treatment further using data from
locations closer and further away from the tip. This may allow more fidelity in the tip region and a smoother
transition of the modelling between inboard and outboard blade sections. As mentioned by Lindenburg,
work remains to be done in predicting the forces near the root of the blade, and the effect of the root vortex.
Problems predicting forces near the root were in fact experienced when using this model. The argument for a
special correction in the region of the blade tip can be equally well made for the root region. Isolating such a
correction from the fundamental stall delay effect is likely, however, to be a considerable challenge. 3-D CFD
simulations, which have previously been shown to be more accurate in this region,14 may help to provide data
to develop simpler correction models in the future.
It is clear from the foregoing discussion that current stall delay correction methods fail to reliably predict
blade loads consistently. This may suggest that the fundamental philosophy that underlies these models may
be flawed and that they may not be correctly representing the flow physics. Limitations in the basic boundary
layer theory on which the models have been developed may further compound this problem. It is suggested
that there is a case to revisit the development of correction methods for wind turbine blade aerodynamics based
on contemporary understanding of the flow. In such a method, the delay in the movement of the separation
point may become a secondary consideration compared with the effect of the spanwise pressure gradient on
the chordwise pressure recovery in the separated flow region. The exact formulation is, as yet, unclear but
will be the focus of further work.

Copyright 2008 John Wiley & Sons, Ltd. Wind Energ 2008; 11:459482
DOI: 10.1002/we
480 S.-P. Breton, F. N. Coton and G. Moe

Conclusion
This paper has presented a study of six models intended to correct 2-D aerodynamic coefficients for stall delay
effects on wind turbine blades.
The models generally produced over-predictions of the loads measured on a wind turbine blade over a range
of operating conditions.
Discrepancies between the measured and predicted loads in the tip region led to the suggestion that a cor-
rection model should include a special treatment of this region.
The model developed by Lindenburg, which includes a special correction in the tip region, was found to
perform best over the range of conditions studied, but still exhibited significant shortcomings.
It was concluded that none of the models studied correctly represented the flow physics and that this was
ultimately responsible for their lack of generality.

Acknowledgements
We would like to thank NREL for providing access to the Unsteady Aerodynamics Experiment wind tunnel
database. One of us (S.-P.B.) would like to thank the Fonds Qubcois de la Recherche sur la Nature et les
Technologies as well as the UNIFOR Foundation for financial assistance in the form of a studentship.

Appendix: Nomenclature
a angle of attack ()
b global pitch angle ()
q local pitch angle ()
f local twist angle ()
finf inflow angle ()
Cl lift coefficient ()
Cl,2-D lift coefficient from 2-D wind tunnel measurements ()
Cl,3-D lift coefficient corrected for 3-D effects ()
Cd drag coefficient ()
Cd,2-D drag coefficient from 2-D wind tunnel measurements ()
Cd,3-D drag coefficient corrected for 3-D effects ()
Cn,2-D normal force coefficient from 2-D wind tunnel measurements ()
Ct,2-D tangential force coefficient from 2-D wind tunnel measurements ()
Cn,3-D normal force coefficient with 3-D correction ()
Ct,3-D tangential force coefficient with 3-D correction ()
Cl difference between the inviscid and corresponding 2-D measured lift coefficient ()
Cd difference between the inviscid and corresponding 2-D measured drag coefficient ()
c chord (m)
r spanwise position (m)
R span (m)
rotational velocity (rpm)
l local speed ratio ()

References
1. Tangler JL. The nebulous art of using wind tunnel data for predicting rotor performance. Wind Energy 2002; 5: 245
257. DOI: 10.1002/we.71.

Copyright 2008 John Wiley & Sons, Ltd. Wind Energ 2008; 11:459482
DOI: 10.1002/we
A Study on Rotational Effects and Stall Delay Models 481

2. Gerber BS, Tangler JL, Duque EPN, Kocurek JD. Peak and post-peak power aerodynamics from phase VI NASA Ames
wind turbine data. Transactions of the ASME 2005; 127: 192199.
3. Himmelskamp H. Profile investigations on a rotating airscrew. MAP Volkenrode Report and Translation No. 832,
1947.
4. McCroskey WJ. Measurements of boundary layer transition, separation and streamline direction on rotating blades.
NASA TN D-6321, 1971.
5. Savino JM, Nyland TW. Wind turbine flow visualisation studies. Technical Report, NASA Lewis Research Center,
Cleveland, OH, 1985.
6. Milborrow DJ. Changes in aerofoil characteristics due to radial flow on rotating blades. Proceedings of the 7th BWEA
Conference, Oxford, UK, 1985; 8593.
7. Madsen H, Christensen H. On the relative importance of rotational, unsteady and three-dimensional effects on the
HAWT rotor aerodynamics. Wind Engineering 1990; 14: 405415.
8. Ronsten G. Static pressure measurements on a rotating and a non-rotating 2.375 m wind turbine blade. Comparison with
2D calculations. Journal of Wind Engineering & Industrial Aerodynamics 1992; 39: 105118.
9. Srensen NN, Michelsen JA, Schreck S. NavierStokes predictions of the NREL phase VI rotor in the NASA Ames
80 ft 120 ft wind tunnel. Wind Energy 2002; 5: 151169. DOI: 10.1002/we.64.
10. Narramore JC, Vermeland R. NavierStokes calculations of inboard stall delay due to rotation. Journal of Aircraft 1992;
29: 7378.
11. Dumitrescu H, Cardos V. Rotational effects on the boundary-layer flow in wind turbines. AIAA Journal 2003; 42:
408411.
12. Schreck SJ, Srensen NN, Robinson, MC. Aerodynamic structures and processes in rotationally augmented flow fields.
Wind Energy 2007; 10: 159178. DOI: 10.1002/we.214.
13. Hand MM, Simms DA, Fingersh LJ, Jager DW, Cotrell JR, Schreck S, Larwood SM. Unsteady aerodynamics
experiment phase VI: wind tunnel test configurations and available data campaigns. NREL/TP500-29955, National
Renewable Energy Laboratory, Golden, CO, 2001.
14. 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/TP500-29494, National Renewable Energy Laboratory, Golden,
CO, 2001.
15. Breton SP, Coton F, Moe G. A Study on different stall delay models using a prescribed wake vortex scheme and NREL
phase VI experiment. 2007 European Wind Energy Conference Proceedings, Milan, Italy, 2007.
16. Snel H, Houwink R, van Bussel GJW, Bruining A. Sectional prediction of 3D effects for stalled flow on rotating blades
and comparison with measurements. 1993 European Community Wind Energy Conference Proceedings, Lbeck-
Travemnde, Germany, 1993; 395399.
17. Chaviaropoulos PK, Hansen MOL. Investigating three-dimensional and rotational effects on wind turbine blades by
means of a quasi-3D NavierStokes solver. Journal of Fluids Engineering 2000; 122: 330336.
18. Raj NV. An improved semi-empirical model for 3-D post-stall effects in horizontal axis wind turbines. Master of
Science Thesis, University of Illinois, Urbana-Champaign, 2000.
19. Bak C, Johansen J, Andersen PB. Three-dimensional corrections of airfoil characteristics based on pressure distributions.
2006 European Wind Energy Conference Proceedings, Athens, Greece, 2006.
20. Lindenburg C. Investigation into rotor blade aerodynamics. ECN-C03-025, Petten, Netherlands, 2003.
21. Corrigan JJ, Schillings JJ. Empirical model for stall delay due to rotation. American Helicopter Society Aeromechanics
Specialists Conference Proceedings, San Francisco, CA, 1994; 8.4-18.4-15.
22. Leishman JG. Principles of Helicopter Aerodynamics. Cambridge University Press: New York, 2006; 589.
23. Bertin JJ, Smith ML. Aerodynamics for Engineers. Prentice-Hall: Upper Saddle River, 1998; 266267.
24. Robison DJ, Coton FN, Galbraith RAMcD, Vezza M. Application of a prescribed wake aerodynamic prediction scheme
to horizontal axis wind turbine in axial flow. Wind Engineering 1995; 19: 4151.
25. Bertin JJ, Smith ML. Aerodynamics for Engineers. Prentice-Hall: Upper Saddle River, 1998; 103104.
26. Bertin JJ, Smith ML. Aerodynamics for Engineers. Prentice-Hall: Upper Saddle River, 1998; 294.
27. Coton FN, Wang T, Galbraith RAMcD. An examination of key aerodynamic modelling issues raised by the NREL blind
comparison. Wind Energy 2002; 5: 199212. DOI: 10.1002/we.58.
28. Sheldahl RE, Klimas PC. Aerodynamic characteristics of seven symmetrical airfoil sections through 180-degree angle
of attack for use in aerodynamic analysis of vertical axis wind turbines. Sandia Report, SAND80-2114 UC-261, Sandia
National Laboratories, Albuquerque, 1981.
29. Hansen MOL. Aerodynamics of Wind Turbines. James and James: London, 2000.
30. Du Z, Selig MS. A 3-D stall-delay model for horizontal axis wind turbine performance prediction. AIAA-98-0021,
1998.
31. Banks WHH, Gadd GE. Delaying effect of rotation on laminar separation. AIAA Journal 1963; 1: 941942.
32. Schepers JG, Brand AJ, Bruining A, Graham JMR, Hand MM, Infield DG, Madsen HA, Paynter RJH, Simms DA.
Final report of IEA Annex XIV: field rotor aerodynamics. ECN Technical Report ECN-C-97-027, 1997.

Copyright 2008 John Wiley & Sons, Ltd. Wind Energ 2008; 11:459482
DOI: 10.1002/we
482 S.-P. Breton, F. N. Coton and G. Moe

33. Lindenburg C. Modelling of rotational augmentation based on engineering considerations and measurements. 2004
European Wind Energy Conference Proceedings, London, UK, 2004.
34. Laino DJ, Hansen CH, Minnema JE. Validation of the AeroDyn subroutines using NREL unsteady aerodynamics
experiment data. Wind Energy 2002; 5: 227244. DOI: 10.1002/we.69.
35. Private e-mail communication with C. Bak, April 2007.
36. Corten GP. Flow separation on wind turbine blades. PhD Thesis, University of Utrecht, 2001.
37. Leishman JG, Beddoes TS. A generalised model for airfoil unsteady aerodynamic behaviour and dynamic stall using
the indicial method. Proceedings of the 42nd Annual Forum of the American Helicopter Society, Washington, DC,
1986; 243265.
38. Tangler JL, Selig MS. An evaluation of an empirical model for stall delay due to rotation for HAWTs. Proceedings
of Windpower 97, Austin, Texas, 1997; 8796.
39. Johansen J, Srensen NN. Aerofoil characteristics from 3D CFD rotor computations. Wind Energy 2004; 7: 283294.
DOI: 10.1002/we.127.
40. Sant T, van Kuik G, van Bussel GJW. Blade pressure measurements on the NREL phase VI rotor using a free wake
vortex model: axial conditions. Wind Energy 2006; 9: 549577. DOI: 10.1002/we.201.
41. Shen WZ, Mikkelsen R, Srensen JN. Tip loss corrections for wind turbine computations. Wind Energy 2005; 8: 457
475. DOI: 10.1002/we.153.
42. Ronsten G, Dahlberg J, Meijer S, Dexin H, Ming C. Pressure measurements on a 5 35 m HAWT in CARDC 12*16 m
wind tunnel compared to theoretical pressure distributions. 1989 European Wind Energy Conference and Exhibition
Proceedings, Glasgow, Scotland, 1989; 729735.
43. Duque EPN, Burklund MD, Johnson W. NavierStokes and comprenhensive analysis performance predictions of
the NREL phase VI experiment. Journal of Solar Energy Engineering 2003; 125: 457467.
44. Laino DJ, Hansen AC. Current efforts toward improved aerodynamic modeling using the AeroDyn subroutines. AIAA-
2004-0826, 2004.
45. van Rooij RPJOM, Schepers JG. The effect of blade geometry on blade stall characteristics. Transactions of the ASME
2005; 127: 496.

Copyright 2008 John Wiley & Sons, Ltd. Wind Energ 2008; 11:459482
DOI: 10.1002/we

You might also like