You are on page 1of 24

WIND ENERGY

Wind Energ., 1, 46±69 (1998)

Review Review of the Present Status of


Article
Rotor Aerodynamics
H. Snel,* Netherlands Energy Research Foundation, Petten, The Netherlands

Key words: A review is presented of the state of the art in wind turbine rotor aerodynamics. Both the
aerodynamics; `engineering methods', used in aeroelastic computer programmes, and more funda-
wind turbines; mental Euler and Navier Stokes are described and reviewed and problem areas in urgent
computational
¯uid dynamics; need of a solution are addressed. An attempt is made to minimise the use of equations in
boundary layers; the main text, to access a wider public. The basic equation systems are treated separately
stall in an Appendix. * c 1998 John Wiley & Sons, Ltd.

Introduction

Fluid mechanics, and speci®cally aerodynamics, is one of the core disciplines of wind energy. It is needed
to describe the wind ®eld from which the rotor extracts energy. It is needed for rotor design, both with
respect to energy yield (performance) and the dynamic loads that develop in interaction with the
elastically deforming structure (aeroelasticity). It is needed to understand and reduce aeroacoustic noise
production and transmission by the rotor blades. It is needed to understand the development of the wake
behind the rotor, which (in wind farm situations) can determine the in¯ow ®eld for the next downstream
rotor.
The basic equations believed to describe the ¯ow of ¯uids (including air) in general, the Navier±Stokes
equations, have been known for a long time. However, their character is such that solutions for the ¯ow
conditions and geometries of interest, even by numerical methods, cannot be obtained as yet. For
practical (industrial) design and analysis purposes a number of approximating models are being used that
permit solutions in a routine fashion. These methods, often referred to as engineering methods, have been
signi®cantly improved over the last decade. At the same time, simpli®ed Navier±Stokes solvers for wind
turbine applications (with limitations such as `Reynolds averaging', turbulence modelling and global
steadiness assumptions) are now being developed and tested in research settings.
This article describes the present state of the art and science of rotor aerodynamics. Both the
developments in engineering models and in Navier±Stokes-like solutions are described, as well as the ever-
growing body of experimental data that is becoming available. Actual problem areas, in urgent need of
solution, are also addressed. Needless to say, the article leans on other recent review articles such as those
by Hansen and Butter®eld,1 Snel and van Holten2 and Sùrensen.3 The present article, however, apart from
being an update in time, is wider in scope than the aforementioned reviews, which are limited to particular
areas.
At attempt is made to minimize the use of equations in the text in order to make the article accessible to
a wider public. In the Appendix an overview is given of the underlying systems of ¯ow equations and
approximations used, including the pertinent physical interpretation. The mathematically more initiated
reader is referred to this.

Rotor Aerodynamics
Rotor aerodynamics refers to the interaction of the wind turbine rotor with the incoming wind. It can be
separated intuitively into a global ¯ow ®eld that extends from far upstream of the turbine to far
*Correspondence to: H. Snel, Netherlands Energy Research Foundation, Petten, The Netherlands. Email: hsnel@ecn.nl

# 1998 John Wiley & Sons, Ltd.


Review of Rotor Aerodynamics 47

downstream, and a local (rotor blade) ¯ow ®eld which refers to the (viscous) ¯ow about the rotor or even
the individual blades. This local ¯ow ®eld can be regarded as a box within (but separated from) the
global ®eld, while boundary conditions (pressures, velocities continuous) on the common boundary unite
the two regions. This separation is used to advantage in the classical blade element momentum (BEM)
method, where a balance is considered between changes in momentum and energy ¯ow rates in the global
part and aerodynamic forces on the blades which are inferred from local ¯ow conditions. The BEM
method, although in a number of details modi®ed from its original form, still forms the backbone of
rotor design and aeroelastic computer programs and will receive due attention. Also in `modern'
computational treatments the methods used for the global ¯ow ®eld usually di€er from those used in the
rotor blade region.
In view of this, the following discussion of rotor aerodynamic methods will be separated into methods
to analyse the global ¯ow ®eld and methods for blade ¯ow analysis. Nevertheless, it is of utmost
importance to realize that these regions are strongly interconnected and cannot quantitatively be
determined without taking into account their interaction:
. the ¯ow in the global region determines the in¯ow conditions for the rotor blades;
. the forces on the blade (possibly expressed as a pressure change) determine the ¯ow in the global region.
Finally, the development of new aerodynamic pro®les speci®cally for use in wind turbine rotor blades
will be reviewed.

The Global Flow Region


The Momentum Method
The classical analysis method for the global ¯ow region is the axial momentum method, in which the
rotor is modelled as an actuator disc. Momentum and energy ¯ows are considered in a control volume
consisting of the stream tube that encloses the actuator disc. This analysis has its roots in propeller theory
as developed by Froude4 and Lanchester.5 Betz6 applied this analysis successfully to the wind turbine
situation, considering the axisymmetric and steady ¯ow about a uniformly loaded actuator disc. The
situation is shown in Figure 1. Although the method and its results are part of the standard knowledge of
the rotor aerodynamicist, some observations will be made here for future reference and background.
For the conditions outlined above, Betz derived the famous limit that bears his name, stating that the
maximum amount of energy extraction from the wind equals the 16/27th part of the kinetic energy
content in the wind, i.e.
C P;max ˆ 16=27 ˆ 059 . . . …1†

where CP denotes the power or performance coecient. Basic to this result are the concepts that there is
no net axial pressure force due to the pressure distribution on the external stream tube (S in Figure 1) and
that there are no external radial forces on the ¯ow. Both concepts are made acceptable by Betz. It must be
noted that according to van Kuik7 the radial force assumption does not hold true owing to an edge
singularity of the actuator disc ¯ow, and that the `real' maximum of the power coecient is expected to
be slightly higher than the Betz limit. Apart from this, ¯ow concentrators such as solid di€users8 or `tip

Figure 1. Basic ¯ow ®eld for global axial momentum analysis

# 1998 John Wiley & Sons, Ltd. Wind Energ., 1, 46±69 (1998)
48 H. Snel

vanes'9 can in principle augment the CP,max value, but up to date this does not seem an economically or
technologically attractive option.
For future reference we state some fundamental results and relations pertaining to the axial momentum
method for a uniformly loaded disc. For a more complete discussion the reader is referred to the excellent
exposition by de Vries.10 First it is noted that upstream of the disc the stream tube enclosing the control
volume will progressively widen on approaching the disc, as the braking e€ect of the disc lowers the
velocity from its original free stream value of V to a value denoted by VD , i.e.
ui
V D ˆ V ÿ ui ˆ …1 ÿ a†V; aˆ …2†
V
where ui is the induced velocity in the rotor plane and the non-dimensional factor a is known as the axial
induction factor. Across the actuator disc a pressure drop occurs to below ambient pressure. In the
downstream part of the ¯ow, the wake region, a gradual pressure recovery occurs until ambient pressure
is reached again. In this process the velocity further decreases until a level of Vw is reached where the
pressures are equalized again. The method assumes no viscous interaction and a completely incom-
pressible ¯ow.
The total power extracted can be written as the product of the mass ¯ow and the di€erence in speci®c
kinetic energy in the free wind and the far wake, i.e.

P ˆ rAR V D 12…V 2 ÿ V 2w † …3†

and likewise as the product of the axial force Dax on the actuator disc and the local ¯ow velocity, i.e.
P ˆ Dax V D …4†

Finally, the axial force can be written as the di€erence in momentum ¯ux in the incoming and outgoing
¯ows:
Dax ˆ rAR V D …V ÿ V w † …5†

Combining (4) and (5) and equating the power extraction as expressed by (3) and (4), it follows that the
velocity VD at the disc equals the algebraic average of the free wind velocity and the far wake velocity, i.e.

V D ˆ 12…V ‡ V w † …6†

or, using the notation of (2),


V w ˆ …1 ÿ 2a†V …7†

Note that basic to this result is the assumption that all the `organized' kinetic energy that is removed from
the ¯own ®eld is extracted at the disc and not, for example, converted into turbulent or recirculation
kinetic energy. Betz's result is straightforwardly obtained from the foregoing equations by maximizing
the power output in terms of the induction factor.
Glauert11 applied the momentum method on an annular level for concentric annuli of radial extension
Dr. This enables the matching of the results of momentum analysis with the blade element properties and
geometry within the speci®c annulus. In this way the method can be converted into a real rotor design and
analysis tool. At the same time an expression for the angular (moment of) momentum balance was added
by Glauert, in which changes in the angular momentum (from the free stream value of zero) are equated
to the torque exerted by the rotor on the air. This involves the introduction of a `tangential' or angular
induction velocity ut . In more recent years, Wilson and Lissaman12 updated the method to include a tip
correction factor (derived by Prandtl) in terms of the induction velocities and in general cast the theory in
a form that lends itself to computer solution; this work forms the basis on which most modern design
tools have been constructed.

# 1998 John Wiley & Sons, Ltd. Wind Energ., 1, 46±69 (1998)
Review of Rotor Aerodynamics 49

The axial momentum method has many de®ciencies both from the theoretical and the practical view-
point. To start with, the application of the momentum and energy relations on an annular level assumes
that the ¯ow in a certain annulus has no relation with the remaining annuli, i.e. annular independence
(this will be referred to as the `annular' momentum method). It is well known that this is not correct.
Goorian13 noted that instead of the momentum analysis on an annular level, a Euler solution of the ¯ow
®eld should be done, of course, after specifying the force distribution on the rotor (actuator) disc. In fact,
in view of the process of vorticity mixing and di€usion in the wake, a complete Navier±Stokes solution is
necessary for a correct description. The next subsection contains a more detailed discussion on this
subject and its present state. Here, it is noted that for axial symmetric ¯ow and rotor loading conditions
that are not too far from uniform, the results from annular momentum theory are in practice quite
acceptable.
However, the annular independence assumption becomes completely inadequate for the case of yaw
misalignment, i.e. when there exists an angle of appreciable magnitude between the wind direction and the
rotor axis. This is a very important condition in practice, since wind direction variations that result from
the 3D character of turbulence are too rapid to be followed by the yawing system of the turbine (either
active or passive), and instantaneous yaw misalignments of 208±308 are frequent. An important part of
fatigue lifetime consumption of rotor blades can be attributed to these conditions, hence a good predic-
tion of the resulting loads is of utmost importance. For that reason, recently a number of improvements
have been proposed and implemented by di€erent groups, as described in References 14 and 15. In these
improved methods a mean speed is calculated in the annuli by applying basic annular momentum theory
to the axial component of the wind speed, but an azimuthal distribution is subsequently applied which
depends on the yaw angle. The form of this azimuthal distribution is taken from an early model by
Glauert11 in some cases, while others base the form on the results of Goankar and Peters.16 These
modi®cations, although certainly not exact, a€ord reasonable comparison with measurements, as
reported in References 15 and 17 and shown in Figure 2 adapted from Reference 17. In this ®gure,
calculated yaw moments for the Tjñreborg 2 MW turbine are compared with the yaw moments derived
from blade root bending measurements. The conditions are indicated in the ®gure. The calculations were
performed with yawed ¯ow models implemented in di€erent aeroelastic computer programs. Models
commonly used before the improvements discussed above resulted in zero-mean yawing moments. The
new models give the correct sign of the (restoring) moment but underestimate the azimuthal variations.
Moreover, it should be stressed that the methods have not been validated for the dynamically changing
yaw angles that occur in practice.

Figure 2. Results of improved yaw models within BEM method

# 1998 John Wiley & Sons, Ltd. Wind Energ., 1, 46±69 (1998)
50 H. Snel

Another basic problem of the momentum method (both global and annular) is that is formulated for
time-independent ¯ow: a possible change with time of the energy or momentum content within the
control volume shown in Figure 1 is not considered. On the contrary, the modelling assumes an instant-
aneous equilibrium between momentum ¯ux di€erence far upstream and downstream and the
aerodynamic forces on the actuator disc (or rotor blades), even for changing conditions. This is known
as the equilibrium wake assumption. In practice, the ¯ow ®eld about the rotor is ever changing in time, as
a result of wind speed and direction changes, wind shear e€ects, blade control and dynamic deformation.
Hence there is no global equilibrium, but an evolving wake situation. These e€ects have been thoroughly
studied both experimentally and theoretically in the framework of the Joule `dynamic in¯ow' projects.
Experiments on the Tjñreborg 2 MW turbine with fast pitch changes revealed large overshoots of the
equilibrium loads. The e€ects can now be modelled by rewriting the equations for the annular momentum
method in the form of dynamic di€erential equations in time instead of the algebraic equilibrium
relations. This type of modelling is presently used in all main design and analysis programs. Descriptions
of various forms of implementation can be found in Reference 17. Sample results of calculations with
these dynamic e€ects and the comparison with measurements are shown in Figure 3, taken from
Reference 17. Again, this ®gure compares results of calculations of blade root bending moments using the
dynamic models with measured bending moments on the Tjñreborg turbine. The measured overshoots
are reasonably well predicted.
Finally, the momentum method breaks down for very high rotor loading, as has been known since the
pioneering work of Betz and Glauert. This condition will typically arise for operation at values of the tip
speed ratio l above 1.3 or 1.4 times the value for which CP,max is attained (usually referred to as ldes). For
the constant rotational speed wind turbine this means that it occurs at low wind speed, much lower than
rated wind speed, at which energy production is relatively low. For a variable speed turbine it may not
occur at all during normal operation. The physics of the ¯ow in this operating mode is such that a
considerable amount of kinetic energy is converted into large-scale turbulent recirculation modes. For
this reason, this condition is known as the `turbulent wake state'. This means that the equality between the
two expressions (3) and (4) for P no longer holds. The usual way of resolving the calculational problem
that results is by using an empirical relation between annular (axial) forces and the induction factor a,
e.g. the so-called Glauert relation. A discussion of di€erent methods used can be found in Reference 18.
It is generally accepted that the current methods tend to give an underestimation of the extracted power at
low wind speeds. No recent work has been done in this respect, although for a ®xed speed wind turbine it

Figure 3. Results of calculations with dynamic in¯ow models compared with measurements

# 1998 John Wiley & Sons, Ltd. Wind Energ., 1, 46±69 (1998)
Review of Rotor Aerodynamics 51

can be estimated that 15%±20% of the yearly energy generation takes place in this condition. Exact values
depend on the matching between design wind speed and annual mean wind speed. If the prediction error
were 20%, then the total error in annual energy yield would be of the order of 4%. There are no important
implications for blade loads, as the phenomenon occurs at low wind speed, i.e. low blade loads. This is
possibly the reason that the problem has not been regarded of high priority.
With the recent adaptations as explained in the preceding paragraphs, the adaptations concerning blade
¯ow (see page 53) and many more that undoubtedly will come, annular blade element momentum theory
can be expected to remain as the aerodynamic model of choice for aeroelastic response calculations for at
least the next 5 years.

Euler and Navier±Stokes Solvers for the Global Flow Field


As mentioned in the preceding subsection, the global ¯ow ®eld should in principle be analysed by more
sophisticated methods in order to overcome the de®ciencies of momentum theory. In this context it must
®rst be stated that because of the ¯ow speeds involved (practically below 25 m s71, which is the cut-out
wind speed for almost all turbines), the assumption of incompressibility is completely justi®ed. With
regard to the blade ¯ow, with blade tip speeds well below 100 m s71, the incompressibility assumption is
still justi®able.
The most rigorous way of analysing the ¯ow ®eld would then be by using the time-dependent incom-
pressible Navier±Stokes equations. However, this leads to the problem of a set of equations that cannot be
solved in a practical sense (see Appendix). The next best solution method is possibly the use of the
Reynolds-averaged Navier±Stokes (RANS) equations with a suitable turbulence model. Although this
statement introduces the problem of deciding which of the many existing turbulence models is the
`suitable' one, the problem can in practice be solved. Still, a precautionary note should be made regarding
the fact that some experts feel that, precisely because of the basic unsteady nature of turbulence and the
need for closure models, RANS methods will not lead to completely satisfactory solutions. These ideas
are expanded somewhat on pages 56 and 58.
One step more down the line of complexity would be the use of the Euler equations instead of the
Navier±Stokes equations. The Euler equations are in fact the non-viscous form of the Navier±Stokes
equations. The processes of creation, di€usion and dissipation of vorticity are not considered by the Euler
equations, but vorticity transport is. There seems to be good justi®cation for this approximation in view of
the large Reynolds numbers involved in the global ¯ow and the fact that there are no solid boundaries in
this region. Hence vorticity will be created only in the local (rotor) ¯ow box owing to the presence of the
blades, or owing to non-uniform disc loading when this part is considered to be an actuator disc. Still, it
may be expected that downstream of the rotor the vorticity created is very much concentrated in the blade
wake and that di€usion may have some e€ects. More seriously, the concentrated vortices that form
especially towards the edge of the rotor (re¯ected as a velocity discontinuity (!) across S in the momentum
method) will create turbulence and mixing with the outer ¯ow. This e€ect is not considered in the
momentum method (although in principle it is possible) and cannot be accounted for in any way by the
Euler equations. The importance of this has recently been investigated, as discussed later in this
subsection.
Euler solutions have been obtained by several authors. First of all, a number of publications have been
devoted to solutions for the axisymmetrical case, where the incoming ¯ow is supposed to be vorticity-free
and only the wake contains vorticity, created at the actuator disc. A thorough discussion can be found in
Reference 19. Although interesting from the theoretical point of view, e.g. to check the seriousness of the
annular independence assumption of the momentum method, this application does not represent a
condition in which momentum theory is very much in error.
A special class of Euler solutions is formed by the so-called vortex wake methods. Usually these
methods are combined with a lifting line or lifting surface representation of the rotor blades. Vorticity
formed at the blades trailing and shed vorticity (see pages 54±56) is convected into the wake with the local
total velocity, calculated as the vectorial sum of the free stream velocity V and the relevant velocities

# 1998 John Wiley & Sons, Ltd. Wind Energ., 1, 46±69 (1998)
52 H. Snel

induced by the existing vorticity elements. These methods are typically unsteady in nature; in the initial
condition the ¯ow is vortex-free and the evolution of the vortex wake is calculated in time. As for vortex
elements, either vortex line elements or vortex particles are used by di€erent groups. Typical applications
with vortex line elements can be found in References 20 and 21 while the group of the National Technical
University of Athens specializes in the vortex particle method, e.g. Reference 22. Line element approaches
have been applied to helicopter aerodynamics also. The aim of calculations with this method is the
prediction of the ¯ow ®eld in the rotor plane, resulting from the undisturbed wind and the induction of all
(or a speci®c part) of the wake vorticity. Among the methods that can be considered as operational at this
moment, two have been exercised to a large extent in Joule projects, i.e. the ROVLM method of the
University of Stuttgart21 and the GENUVP method of the National Technical University of Athens.22
The important advantage of the vortex wake methods is that they lend themselves straightforwardly to
the calculation of unsteady situations, in the in¯ow or with regard to the rotor blade pitch angles, and can
be applied straightforwardly to the yaw misalignment situation. However, a theoretical problem exists
with regard to the stability of the calculations, especially for the vortex line element case. This problem is
basic to all curvilinear vortex elements and their discretized version (rectilinear elements that connect to
each other at an angle di€erent from 1808): the self-induced velocity `has a logarithmic singularity.
Physically, the vorticity is not concentrated in (singular) lines of zero thickness, so the problem is only a
mathematical one. The consequence is that without precautionary measures the method does not
converge when the element size is reduced to ever smaller values, since the `collocation' points where
velocities are calculated (usually the midpoints of vortex line elements) come too close to neighbouring
vortex line elements. The same problem occurs when during the convective process being simulated a
vortex collocation point moves too close to a line element and the point is ejected at very high speed. In
that case the result becomes something like `vortex spaghetti' and the computation breaks down. One way
to overcome this problem is to prescribe either a cut-o€ length (no induction for separations smaller than
the cut-o€ length) or a viscous vortex core with regular velocity. However, the solution then becomes a
function of the choice of this cut-o€ or viscous core length, which certainly is not a part of the Euler
solution. Fortunately, it can be made probable by simple model studies23 that this dependence is very
small indeed, as long as the cut-o€ distance has a value that can be estimated on the grounds of either
empirical information (e.g. Reference 24 concerning vortex core size) or the action of viscosity.
A practical disadvantage of the free vortex wake methods is the very large amount of computer CPU
time needed for a calculation. In order to obtain a realistic ¯ow situation in the rotor plane (needed for the
calculation of the blade ¯ow), the calculated wake must extend to at least two rotor diameters behind the
rotor plane. This represents some 15 rotor revolutions and a large number of vortex elements that have to
be traced. Calculation times involved in free wake calculations are some tens of thousands times larger
than those needed for momentum methods. Because of this, many so-called prescribed wake methods
have been developed, in which the geometry of the wake is either ®xed or described by a few parameters
which determine its shape. Hybrid methods have also been developed, in which the near wake (e.g. one
rotor diameter downstream) is treated as a free wake and the rest as prescribed wake. The ROVLM
program of the University of Stuttgart o€ers these possibilities and has checked the results against
calculations with a completely free wake. It was reported by Bareiû et al.21 that a 75% reduction of the
computer time can be obtained by accepting 5% di€erences with the free wake calculations. It should be
noted that for the hybrid wake the downstream prescribed wake part does have an in¯uence on the
development of the near (free) wake.
Although much computational e€ort can be saved by prescribed wake methods, the advantages of such
a computation over the (annular) momentum method are not so clear. In fact, annular independence is
now replaced by a `prescribed' dependence, without certainty about the correctness of the prescription
other than the comparison with free wake calculations for some conditions.
Finally, yet another approach which e€ectively is a Euler solver is represented by the so-called
asymptotic acceleration potential method. This method has been developed and used mainly at Delft
University of Technology. It was ®rst formulated25 in order to extend Prandtl's lifting line theory to the

# 1998 John Wiley & Sons, Ltd. Wind Energ., 1, 46±69 (1998)
Review of Rotor Aerodynamics 53

case of rotating wings with unsteady ¯ow for helicopter applications. In Prandtl's theory the induction
due to the trailing vorticity system from a stationary 3D wing is used to adapt the local in¯ow velocity
®eld. The method was expanded and implemented by van Bussel26 for application to wind turbines. A
cautionary note should be made that the method assumes small perturbations of the main ¯ow, which is
justi®ed in helicopter applications but much more doubtful for wind turbines. Because of its linearity, the
method is much more ecient in computer time than straightforward vortex wake methods but still
belongs to a category that does not compete with momentum methods. Like vortex wake methods, its
applicability is mainly in examining ¯ow ®elds and inferring improvements to momentum methods. It has
been used as such in the Joule `dynamic in¯ow' projects, and a recent example for yawed ¯ow can be
found in Reference 27.
Navier±Stokes solutions for the global ¯ow ®eld are not often used but can yield interesting informa-
tion. For example, Madsen28 applied a Navier±Stokes solver with a k±e turbulence model to axially
symmetric ¯ow across an actuator disc. For a disc loading corresponding to the maximum energy
extraction condition, (i.e. a ˆ 13; Dax ˆ 49rV 2 AR ) the solutions shown very little e€ect of turbulent wake
mixing on the conditions in the rotor plane. The most interesting results are obtained for high rotor
loading, corresponding to Dax 4 12rV 2 AR , where the momentum theory breaks down and Glauert's
correction is often used. For this case the Navier±Stokes solution does show a remarkable in¯uence of the
turbulence mixing. For a turbulence mixing of 10% the numerical solution follows the Glauert correction
reasonably well for values of the axial force coecient CD,ax up to 1.15, but beyond that value the
induction obtained by the Navier±Stokes solutions is appreciably higher than that indicated by Glauert.
Hence turbulent mixing of the wake seems to be important, but only for high axial loading.
In e€ect, the important use of Euler methods or Navier±Stokes solutions at this stage of development is
in providing information on which to base corrections to the momentum method which is used in
aeroelastic design codes. This is especially true for situations such as yawed ¯ow or high rotor loading.
Applied in this manner and used with sucient precaution, the methods have the (far) future potential to
be used as a substitute for experimental work. Of course it must be stressed that before this can be the
case, the methods themselves must be subjected to a long process of validation against results from
measurements (with wind tunnel quality and of suciently high Reynolds number) in order to be
trustworthy.

Blade Flow Analysis


The ¯ow conditions just upstream of the rotor plane, as determined from the analysis of the global ¯ow
®eld, form the `in¯ow' condition for the rotor blades and determine the aerodynamic forces on the
blades. It is repeated that the magnitude of these forces (e.g. in terms of pressure drop across an actuator
disc with non-uniform loading) in turn forms the needed input for the analysis of the global ¯ow ®eld.
The actual matching of these ¯ow regions depends very much on the method of analysis that is chosen.

2D Blade Element Theory


Blade element theory is the method of choice for the analysis of blade properties when the annular
momentum method is used for the global ¯ow region. In fact, the application of the momentum (axial
and angular) equations results in analytical relations between the induced velocity components and the
aerodynamic forces exerted by the elements of the blade in this region. In 2D blade element theory, it is
assumed that these forces are equal to the forces on the same aerodynamic pro®le taken from two-
dimensional wind tunnel aerofoil tests. Hence the force characteristics in terms of the non-dimensional
lift and drag coecients Ci(a) and Cd(a) can be taken from wind tunnel measurements, as long as the
Reynolds number is roughly the same in the measurements and in the turbine situation. This approach is
based on Prandtl's slender wing (lifting line) approximation, in which, for an aerodynamic wing with a
large span-to-chord ratio, the forces on an element are taken equal to the 2D forces for an equivalent
angle of attack that are formed by the mean ¯ow plus the velocities induced by the 3D trailing vortex

# 1998 John Wiley & Sons, Ltd. Wind Energ., 1, 46±69 (1998)
54 H. Snel

system. In the case of the wind turbine this 3D induction is the result of the helical trailing vortex system
in the rotor wake and can be equated to the induction velocities of momentum theory. Relating the
aerodynamic forces (calculated in the manner) in the annulus to the momentum loss in the same, a closed
set of non-linear algebraic equations is obtained which can be solved numerically. This is the feature that
makes the combination of momentum theory and 2D blade element theory computationally very
ecient.
In the state-of-the-art models used for aeroelastic analysis, a number of corrections are applied to this
strictly 2D method, such as the so-called tip correction and, more recently, a correction for 3D e€ects in
stall and models for dynamic stall. The correction known as the `tip' correction accounts for the fact that
induction is not uniform over the annulus under consideration owing to the ®nite number of blades and
the resulting non-uniform vorticity distribution in the wake. In fact, the tip correction is a non-uniformity
correction. There are several ways in which this correction can be applied; see for instance Reference 10
for a discussion of the various possibilities. The most common implementation is the one introduced by
Wilson and Lissaman,12 based on an analysis by Prandtl. This correction, which depends on the tip speed
ratio, the number of blades and the radial position, is only important in the blade tip region and e€ectively
changes the local angle of attack, which tends to the zero-life angle approaching the tip.
The 3D correction for stall is essentially unlike the tip correction in the sense that it is a correction to the
2D (wind tunnel determined) aerofoil characteristics, mainly to account for the e€ects that radial ¯ow has
in a rotating system. This e€ect is completely di€erent from that in a non-rotating system, since Coriolis
forces are introduced that act as additional pressure gradients on the ¯ow about the blade section. This
e€ect is explained in more detail on page 57. It has been in¯uential in improving BEM results, compared
with measurements, for stalled ¯ow.
Finally, when analysing an unsteady situation (turbulent wind, dynamically responding blades, etc.),
the measured steady Cl(a) and Cd(a) data are usually replaced by a model for dynamic lift and drag
coecients. This is especially important in stall, as important hysteresis loops form. In fact, generally
speaking, the dynamic stall loops of the lift coecient contribute to stable ¯ap (out-of-plane rotor blade
bending) dynamics, whereas the use of steady coecients in stall sometimes leads to unstable results of the
computation. Stable ¯ap behaviour is observed experimentally. There are a relatively large number of so-
called engineering models in use for the simulation of dynamic stall phenomena. These models have the
form of ordinary di€erential equations in time and are completely empirical in nature. A recent overview
of methods presently in use can be found in Reference 29. There is still considerable ongoing activity in
this ®eld, among others towards the modelling of the self-excited vortex shedding that causes higher-
frequency unsteadiness in stall. Much of this is based on the approach pioneered by Truong.30
Accurate predictions of 3D stall delay and of dynamic stall characteristics are still the most challenging
problems for aerodynamic analysis. Although considerable progress has been made, both in under-
standing and in modelling, the results are not yet suciently accurate when compared with measurements.
The prediction of the maximum power for a stall-controlled turbine is within 15% accurate, while power
curve guarantees must be given by manufacturers of 5% accuracy. Also, there are still serious discrep-
ancies in the simulation of some dynamic problems in stall, especially the in-plane dynamics of large
blades. This will be discussed on page 58.

Lifting Line and Lifting Surface Theory


Lifting line or lifting surface theory is usually applied to blade ¯ow analysis when vortex wake methods
are used for the global ¯ow. In the lifting line method the rotor blades are modelled as `bound vortex
lines', i.e. as a line (geometrically coinciding with the blade quarter-chord line) along which a vortex
strength is de®ned that will vary with radial (span wise) position. The local vortex strength G is related to
the local angle of attack, often through a (potential ¯ow) relation between circulation and lift:

cC l …a†V ef
Gˆ …8†
2

# 1998 John Wiley & Sons, Ltd. Wind Energ., 1, 46±69 (1998)
Review of Rotor Aerodynamics 55

where c is the local chord, Cl is the lift coecient as a function of the angle of attack a, and Vef is the
e€ective in¯ow velocity. Vef is vectorially composed of the wind speed, the relative speed Or due to
rotation and the induction due to the wake (see next paragraphs). The local angle of attack is the angle
between the blade chord and the Vef vector. When applying (8) and using Cl(a) from wind tunnel
measurements, the lifting line method is equivalent to the 2D blade element method as far as blade ¯ow
analysis is concerned.
When combined with vortex wake methods, there will be vorticity `trailing' from the vortex line, in
principle in the form of a vortex sheet, with local strength equal to the spanwise derivative of the `bound'
lifting vortex line strength G. In the computational methods this is sheet discretized, e.g. in line element
vortices for the free vortex wake described in Section 2.1.2 The trailing vortex elements which have an
initial direction (of the rotation vector) normal to the blade axis are convected into the wake and
contribute to the induction in the rotor plane.
In the case of unsteady conditions, apart from trailing vorticity associated with changes in the
circulation along the lifting line, so-called `shed vorticity' will be produced, associated with changes in the
bound vortex strength with time. This is comparable with the well-known starting vortex that is created
when an aircraft wing builds up its lift. Shed vorticity has an initial direction parallel to the lifting line. Its
contribution to `induced velocities' must be very carefully considered. In fact, in blade element theory
(or lifting line theory) the induced velocities are used to modify the free stream velocity magnitude and
direction in such a way that the ¯ow situation becomes comparable with the 2D situation and hence 2D
theory or experiments can be applied locally to obtain sectional forces. Consequently, only the vorticity of
3D type must be used to determine the induced velocity. This includes all trailing vorticity, but not all of
the shed vorticity. Indeed, in the unsteady 2D situation there will be a shed vorticity system also, but this is
included in the 2D theory or experiment. Hence, for this system, only the di€erence between the 2D and
3D situations should be used in accounting for the induced velocity. This is mathematically formulated in
Reference 25. In the developing wake this is very dicult to trace and it is not always clear from the
publications in this ®eld whether the implementations are consistent in this manner.
In lifting surface theory the blade is represented in more detail. Instead of a bound vortex line, a
distribution of vortex elements over a surface, usually the blade's camber surface is employed. In this case
no 2D information is used. Instead, for each vortex element a boundary condition is used for ¯ow
alignment with the surface. All vorticity, wake and bound, must be used to ®nd the ¯ow velocity vector,
and a system of equations in terms of the bound vortex strength of the surface elements ensues. The
Stuttgart ROVLM program makes use of the lifting surface method. Obviously, this method is much
more demanding in computation time than the lifting line method.
In the lifting surface representation the thickness of the blade sections is ignored. This can be amended
by representing the real blade geometry in the calculations and by the use of a surface singularity
distribution to resolve the inviscid and incompressible ¯ow equations. Usually, source distributions on the
outer blade surface are used together with a vorticity distribution along the camber surface. This method,
usually referred to as the `panel' method, has been used quite extensively in aircraft aerodynamics, but
only very few publications for wind turbine applications can be found. Reference 31 is one of the few
examples. From the results it may be concluded that the added computational complexity is in no way
compensated by the improved quality of the results.
The main advantage of the lifting surface and panel methods is that they are 3D in character. However,
3D e€ects are less important in attached ¯ow. For separated ¯ow the fact that lifting surface and panel
methods are based on the inviscid ¯ow equations turns into a signi®cant disadvantage. In this respect the
lifting line method has a relative advantage in that it allows a simple introduction of separated ¯ow
characteristics by using empirical Cl(a) data in (8), possibly expanded with 3D stall delay models and
dynamic stall models. Another disadvantage of the inviscid ¯ow theory used in the lifting surface and
panel methods is that pro®le drag is not predicted, even in attached ¯ow. Still, this quantity is very
important for the prediction of the turbine's main shaft torque and power. In order to get acceptable
predictions of these quantities in computer codes based on lifting line methods, the pro®le drag must be

# 1998 John Wiley & Sons, Ltd. Wind Energ., 1, 46±69 (1998)
56 H. Snel

added `arti®cially' to the load calculations, based on an estimated value of the angle of attack or based on
sophisticated boundary layer calculations.
The only way in which more sophisticated blade ¯ow models can contribute notably to the under-
standing of real phenomena and improvement of the accuracy of design methods is by including the
viscous e€ects, as discussed in the next subsections.

General Aspects for Viscous Flow Modelling


Before entering into a more detailed discussion of the present possibilities of viscous ¯ow modelling, it is
important to realize the fundamental problems inherent to this. Basically, these problems can be related
to the fact that the ¯ow about the blade will, at least partly, be turbulent in nature. In laminar ¯ow at low
Reynolds number the viscous ¯ow equations can be solved numerically without too much diculty.
However, owing to the large Reynolds numbers involved (a typical value is Re ˆ 5  106 for a megawatt-
size wind turbine), the adverse pressure gradient on the blade suction side, blade roughness, radial ¯ow
due to yaw error and in¯ow turbulence (1%±2% turbulence intensity is typical for the blade ¯ow
problem), small-scale turbulence will develop in the shear layer on the blade surface.
It is of importance to know at what location the transition from laminar to turbulent ¯ow will occur,
but in view of the large number of parameters involved, this is next to impossible. In fact, Madsen et al.32
suggest that the actual problem of double stall* can be traced to di€erent transition locations, in¯uenced
perhaps by the above parameters. In general the aerodynamic characteristics of the blade will be
in¯uenced by the location of transition. In current practice of the application of boundary layer theory the
general procedure is either to have the user ®x the transition point or to let the computer program decide
where transition takes place, the latter based on the value of some boundary layer speci®c quantities. For
2D ¯ow the so-called spatial ampli®cation theory of van Ingen33 and Smith34 is often used, but the
extension to 3D ¯ows is not well established. Moreover, even in 2D it is questionable to what extent the
above parameters can be taken into account for the determination of the transition location. The situation
is worse in the case of Navier±Stokes solvers, since many commercial packages do not even permit
transition but perform only completely laminar or completely turbulent calculations.
Another fundamental problem is turbulence modelling. For high-Reynolds-number ¯ows the only
practical way of treating turbulence is through a procedure known as `Reynolds averaging' (see
Appendix), which leads to a larger number of unknowns (the so-called Reynolds stresses) than equations.
In order to resolve this diculty, closure relations must be introduced, the so-called turbulence model. A
large number of such models exist in the literature, which are all semiempirical. The most widely used
turbulence model is the so-called k±e model,35 in which k represents the turbulent kinetic energy and e its
dissipation rate. This model is usually applied together with a version of the turbulent wall law. This is
convenient in the sense that the region directly adjacent to the solid surface, where the largest gradients
occur, is solved in an analytical manner. However, it is known that this model is not suciently accurate
for ¯ows with important adverse pressure gradients and with large separated regions, producing too
high values of the shear stress. Possibly the k±o model developed by Wilcox36 together with an adaptation
proposed by Menter37 improves the situation. Indications to this e€ect can be found in Reference 38.

3D Boundary Layer Methods


A relatively ecient way of introducing the e€ects of viscosity into calculations for high-Reynolds-
number ¯ows is by separating the ¯ow domain into an inviscid potential ¯ow part and a thin viscous
boundary layer on the surface of the blade. The equations to be analysed in each of these regions are far
simpler than the complete Navier±Stokes equations and for this reason it has been the method of choice
for aircraft wing aerodynamics (see Appendix). For the outer region a Euler or even a potential ¯ow
solver can be used. The pressure ®eld determined with this method is used as input to the boundary layer

*Double or multiple stall is the phenomenon in which a turbine can operate at distinct power levels under e€ectively the same external conditions. The
cause of this phenomenon is not known with certainty.

# 1998 John Wiley & Sons, Ltd. Wind Energ., 1, 46±69 (1998)
Review of Rotor Aerodynamics 57

equations. In the boundary layer approximation the spatial pressure derivatives in the two surface
directions are used as driving forces, while variations across the layer (normal to the surface) can be
neglected. If desired, a quantity known as the displacement thickness can be obtained from the boundary
layer solution and used to obtain a correction on the potential ¯ow solution of the pressure ®eld.
A basic problem with the boundary layer equations when used with a prescribed pressure ®eld is the fact
that the solution becomes singular at a ¯ow separation. This causes numerical divergence problems with
the corresponding computer programs. The problem can be solved by using what is called a strong
interaction technique, in which the boundary layer equations and the outer ¯ow equations are solved
simultaneously or in a more sophisticated iterative manner than described above. Details about this
process for an application to rotor blades can be found in the thesis of Sùrensen,39 who pioneered the
application of these techniques in a complete 3D fashion for wind energy use, obtaining very interesting
results clarifying the importance of radial ¯ows in the rotating case. The computational results showed the
development of radial ¯ow especially when separation occurs. Radial transport of retarded boundary
layer material towards the tip (in the separated case travelling essentially with the chordwise blade
velocity) has the e€ect of a favourable (i.e. accelerating) chordwise pressure gradient. The result is that the
separated wake for the rotating blade is thinner than for the corresponding 2D case, which means higher
lift and lower drag forces. The e€ects are especially noticeable close to the blade root, decreasing in
magnitude for stations closer to the tip. Recently, boundary layer measurements (for a Reynolds number
of 100 000) with LDA techniques showed radial ¯ow components of the same order of magnitude as the
(reversed) chordwise velocities in separation.40 Calculations with the boundary layer approximation, even
in strong interaction with the outer ¯ow, usually break down in deep stall.

Quasi-3D Boundary Layer Methods, the Case of Stall


Although 3D boundary layer methods are an order of magnitude less demanding in computational time
compared with Navier±Stokes solutions, they are still of a magnitude that precludes their use outside the
research environments. Also, for the design of aerofoils for use in stall-controlled wind turbines there is a
need for a less expensive and quicker method. This was made possible by the development of a quasi-3D
boundary layer method, which took place mainly in the Netherlands. Based on physical insight obtained
from the complete 3D boundary layer calculation results, Snel41 presented an order-of-magnitude
analysis of the di€erent terms in the 3D boundary layer equations for a rotating slender blade in terms of
the local chord-to-radius ratio c/r. It was shown that for the case of separated ¯ow the radial `parts' of the
convective acceleration are of order (c/r)2/3 compared with the main terms (relatively of order one). In the
chordwise boundary layer momentum equation the Coriolis force term remains O(1) and contains the
radial velocity component. Hence the radial momentum equation has to be resolved also. Neglecting all
terms of O(c/r)2/3 and smaller, both the momentum equations and the continuity equation do not contain
radial `convective' derivatives and can be solved in a 2D (stripwise) fashion. This is the crucial advantage
of the method and enables the solution of the equations to be done on a fast personal computer in a
tolerable turnaround time.
The method has been implemented in (originally 2D) boundary layer strong interaction codes, namely
in the NLR ULTRAN-V code42 and later in the commercial XFOIL code. In comparison with measured
results, the calculated results show qualitative agreement, but they need improvement in the quantitative
sense. However, used with judgement, the method o€ers a tool for pro®le design and analysis that can be
used on a PC.
An important drawback of the method is its application in the blade root region, where the radial ¯ow
e€ects are important but where c/r has a value of about 0.25. Neglecting terms of order (c/r)2/3 then
becomes doubtful. Another drawback of boundary layer methods in general is that the conditions under
which the boundary layer approximation is valid (boundary layer thickness small compared with the
chord) perhaps include the initial stall regime, but certainly not deep stall, where the wake thickness is of
the order of the chord length. In fact, this limits the applicability to an approximately 208 angle of attack.
Stall-controlled (or partial span pitch-controlled) wind turbines will operate with angles of attack close to

# 1998 John Wiley & Sons, Ltd. Wind Energ., 1, 46±69 (1998)
58 H. Snel

358 or 408 in the blade root sections at high wind speed. For these conditions a complete Navier±Stokes
solution is the only possibility.

Navier±Stokes Solutions
Navier±Stokes solutions, both 2D and 3D are a current theme for many research groups in the ®eld of
wind energy. It must be stressed, however, that all published work considers the Reynolds-averaged
Navier±Stokes (RANS) equations, which need an empirical turbulence closure model. This is the only
practical possibility for external high-Reynolds-number ¯ows at present. There are some important
problems regarding the usefulness of the method. One fundamental question regards the applicability of
RANS, which averages the turbulent ¯uctuations (by either time or ensemble averages) of turbulent ¯ow
®elds which are unsteady by their very nature into a ¯ow ®eld of statistical averages. A growing number
of experts seem to believe that there is no way around direct numerical solutions (DNSs) of the Navier±
Stokes equations or at least large-eddy simulations (LESs), which do represent (at the larger scale) the
unsteadiness in a direct manner. If this opinion is shown to be correct, then useful Navier±Stokes
solutions are even much further away, as such solutions for 3D ¯ows around blades at high Reynolds
numbers are very far removed from what is possible at this moment.
Even with the RANS method, practical problems regarding transition and turbulence modelling exist
(see page 56). Wolfe and Ochs43 report their use of a commercial CFD code (ACE) to compute the ¯ow
about an S809 pro®le and an NACA 0012 pro®le, comparing the computed results with measurements of
Ohio State University.44 They observe that no commercial Navier±Stokes code includes a transition
criterion. Upon their speci®c request, a possibility of modelling the e€ects of transition was built into the
code, and only with this aid was a good solution obtained for attached ¯ow conditions. A distinct
disadvantage is that a change in the prescribed transition location results in the necessity of constructing a
new grid. A dependable and practical solution of the transition prediction problem is urgent. Also they
argue that the less than satisfactory result for deep stall conditions, and indeed the poor prediction of the
maximum life coecient for the S809 aerofoil, is possibly caused by the de®ciency of the k±e model for
the condition of stalled ¯ow.
Chaviaropoulos45 describes recent work at CRES (Greece) in the realm of 2D unsteady and quasi-3D
Navier±Stokes modelling. Here also the problem regarding the turbulence modelling is noted. The code
developed at CRES (research code) does allow the possibility to use ®xed or free (determined during the
computation) transition, but di€erences arising from this are not discussed. For the quasi-3D modelling,
the expansion in terms of c/r, as discussed on page 57, is used. Qualitatively, the e€ect of increased lift in
separated ¯ow is reproduced.
Finally, some of the work by Sùrensen and colleagues at the Technical University of Denmark (TUDk)
in co-operation with Risù must be noted. Within this co-operation the EllipSys3D steady Navier±Stokes
solver was developed. In Reference 46 this program is applied in a complete CFD approach for the ¯ow
about a wind turbine rotor. The global ¯ow ®eld is resolved by an axisymmetric Euler solver, with the
rotor represented by an actuator disc. The rotor ¯ow is solved by the use of EllipSys3D. In fact, the blade
¯ow is ®rst solved as if no induction occurs. Next the blade forces calculated are introduced as pressure
jumps across the actuator disc. The resulting induction is calculated and with this the blade in¯ow is
reformulated. This process is repeated until convergence is reached. With regard to turbulence modelling,
the Menter37 adaptation of the k±o model is used with good results.
When arguing that 3D Navier±Stokes solutions are important especially in the case of stalled ¯ow, it
should be stressed that up to date no non-stationary time-realistic 3D Navier±Stokes solutions for wind
turbine blades have been reported. Although correct modelling of steady stall (or rather time-averaged
stall) is important for the prediction of power production, it is well known from wind tunnel measure-
ments that stall is intrinsically unsteady in character, even in stationary external conditions, and to an
even much larger extent in the ever-changing conditions of the real wind turbine. This is of utmost
importance for load ¯uctuations and even for the aeroelastic stability of the blade. With regard to out-of-
plane (¯apwise) blade oscillations, it is now generally accepted that dynamic stall phenomena are

# 1998 John Wiley & Sons, Ltd. Wind Energ., 1, 46±69 (1998)
Review of Rotor Aerodynamics 59

responsible for an e€ective damping, whereas the stationary Cl ±a curve would give rise to negative
damping. State-of-the-art `engineering' dynamic stall models are able to model this relatively well.
However, for in-plane (lead±lag) blade oscillations, aerodynamic damping may not be improved by
dynamics. In fact, recent experiences with large stall-controlled rotor blades have shown extreme in-plane
load ¯uctuations (up to ®ve times the gravitational amplitude), which are held responsible for some recent
blade ruptures. Although the exact cause is not known and it is hypothesized that part of the problem is of
a structural dynamics nature, aerodynamic damping or lack of it must play a role. This underlines the
importance of improved understanding and modelling. It is ®nally observed that even blades of pitch-
controlled turbines may experience dynamic stall phenomena through wind gusts around the rated wind
speed.

Available Experimental Data and Need for Controlled Measurements


The validity of analysis and calculation methods can only be shown by comparison with experimental
results. At the same time, these results add to physical insight that is essential to further improve the
analysis methods. Although experiments on rotating blades are essential, non-rotating measurements can
be helpful in comparing results with rotating data and establishing di€erences, to separate rotational
e€ects from other 3D e€ects.
Early ®eld tests were done by Hales47 on a small rotor mounted behind a car, both in yaw and in stall.
The pressure distribution was measured through pressure taps on a few sections. These were the ®rst tests
that showed the particular behaviour of stall on a rotating blade in detail. Still, the Reynolds number
(around 3  105) of these data was relatively low. Nevertheless, it marked the beginning of a series of ®eld
tests, also for larger machines.
Recently, quite a number of institutes have been engaged in aerodynamic ®eld tests on ®ve di€erent
rotors of di€erent sizes. Results of these are now available in a database within an IEA activity known as
Annex XIV.48 The rotor diameters of the turbines used range from 10 m (NREL test facility and Delft
University test turbine) to 27.5 m (ECN HAT 25 test facility). On the largest of these machines, Reynolds
numbers of typically 2.5  106 can be attained, which are suciently high to make the results of direct
interest. The objective of the di€erent test programs was to measure aerodynamic forces on the blades,
which in almost all cases was done with the aid of pressure taps on a number of sections. Only the Risù
turbine measures the forces by way of three-component force balances in which three di€erent sections of
small radial extent are suspended. A detailed description of the test facilities, the data acquisition systems
and the measured data is contained in Reference 48. Special attention was given to stalled operation and
yaw misalignment, since the most urgent unresolved problems lie in this area. The typical triangular
pressure distribution for rotating blade stall at inboard blade sections, as opposed to the ¯at distribution
in the separated region that is measured in the wind tunnel, was measured for all rotors.
For yaw conditions, much of the information present in the database can be used undoubtedly to verify
yaw models in wind turbine response programs. This not only concerns the blade aerodynamic forces, but
also measured load signals on the blade. Much of this type of analysis remains to be done. Present plans
are to establish a new IEA Annex in which the `owners' of the database will make structured use of the
wealth of data available there.
Nevertheless, it must be realized that the comparison of ®eld measurements with computational model
data is a dicult task. The largest problem is that the in¯ow in the rotor plane cannot be determined in a
deterministic sense, since the only information is given by the registration of wind speed and direction at
(at best) a number of stations at some distance (hopefully upstream) from the rotor. Also, wind speed and
direction will be non-uniform over the rotor and the distribution will vary stochastically. Basically, only
statistical data values can be obtained with some degree of accuracy, both for wind speed and wind
direction.
A di€erent problem underlies the determination of the angle of attack pertaining to the measured
pressure distribution on a section. Even if the in¯ow were completely determined, still the angle of attack

# 1998 John Wiley & Sons, Ltd. Wind Energ., 1, 46±69 (1998)
60 H. Snel

as de®ned in the modelling includes the induction of the wake vorticity (or the relevant part of it). More
information regarding the in¯ow conditions and the angle of attack could be obtained if the ¯ow ®eld in
the rotor plane could be measured. This, however, is not a practical proposition for ®eld measurements.
Moreover, even these data should be treated with considerable care, as the bound blade vorticity and the
2D part of the shed vorticity in¯uence the measurements but should not be included in the de®nition of
the angle of attack. In the end the diculties are related to the fact that the angle of attack is a 2D concept
that is de®ned in a wind tunnel environment. Its generalization to the case of 3D rotating blades is by no
means trivial, and the practical translation to measured quantities is a matter of further research.
Notwithstanding the enormous value of ®eld measurements at representative Reynolds numbers, the
above observations state a clear case for sophisticated wind tunnel measurements. At present, valuable
pressure data on rotating blades exist for small Reynolds numbers (between 0.5  106 and 0.7  106) in the
form of FFA measurements on a two-bladed rotor in the Chinese CARDC tunnel. These data, however,
were obtained with a slow measurement system only capable of taking data averaged over a rotation.
Nevertheless, these data have been used extensively by many groups for validation of stall models. Flow
®eld information has been obtained in the past mainly in the Delft University open jet wind tunnel, again
for very low Reynolds numbers. This information has been and will be valuable in the improvement of
yaw modelling.
Another type of measurement that is needed is that of unsteady stall characteristics, ®rst in a 2D tunnel
environment, for the aerofoils to be used in wind turbines and for the typical values of reduced frequency
for this application (below 0.1). One step in this direction was set by the very useful measurements at
Ohio State University50 which were used extensively in the Joule `dynamic stall and 3D e€ects' project.
Again, extension of 2D knowledge to the case of dynamic stall on rotating blades is very dicult. In the
®eld measurements discussed above, dynamic stall conditions have certainly been encountered, but the
measured signals are very much in¯uenced by stochastic wind speed and direction changes, on which
only some statistical information can be obtained. Perhaps with the aid of careful ®ltering, and by
concentrating on regions such as blade tower passing, some information can be obtained from the
existing measurements. In fact, work in this direction is done within the Joule Stallvib project and will no
doubt be reported in the near future. However, it is likely that only controlled wind tunnel measurements
can give sucient detail in information. Such measurements should be done on a suciently large rotor
to obtain Reynolds numbers of at least approaching 106, since for the typical reduced frequencies of wind
turbine rotor blades, viscous e€ects in the boundary layer dominate the dynamics and hence Reynolds
e€ects must be expected. Also, the model should be equipped with fast pressure sensors in order to
capture fast changes in pressure distributions. The importance of this knowledge was discussed on pages
58±59.
If ¯ow ®eld quantities (velocities in or directly behind the rotor plane) could also be measured, together
with the known and controllable in¯ow, de®nitive information could be obtained on many aspects. Also,
such measurements should yield data against which to evaluate the results of 3D Navier±Stokes solvers
that are becoming available. Plans to do such measurements are in development, both in the United States
(NREL) and in the context of European collaboration. It is this author's conviction that this type of
measurement presents the only way to remove, at least to some extent, the main remaining areas of
uncertainty in aerodynamic knowledge.

Aerofoil Design
A speci®c item in the application of aerodynamics is the development of new aerofoil sections speci®cally
designed for use in wind turbine rotors. In the early stages of development the pro®les used more often
were those of the familiar NACA and NASA section families, also known as general aviation pro®les. In
fact, the NACA 44xx, NACA 230xx and NASA LS1-mod sections were among the popular aerofoils in
the very beginning, later being replaced by the NACA 63 AND NACA 64 families, especially for stall-
controlled rotors. The latter are still being used in the outboard part of many blades, but the continuous

# 1998 John Wiley & Sons, Ltd. Wind Energ., 1, 46±69 (1998)
Review of Rotor Aerodynamics 61

quest for lighter blades has led to the introduction of thicker aerofoils in the root section. This has been
one motivation for the development of speci®c wind turbine aerofoils. Another important driving force
has been the desire to use aerofoils with very low susceptibility to the e€ects of dirt accumulation and/or
fabrication irregularities. This aspect is especially important for the operation of wind turbines in the
desert-like environment of Californian wind farms.
Work on dirt-insensitive aerofoils started in the United States, where Tangler and Somers51 have been
active in the development of the S pro®le family, both thick and thin. These pro®les have been utilized
exclusively in the United States. In Europe the work of BjoÈrck at FFA52 must be mentioned (FFA-W
family) and the work of Timmer and van Rooij at Delft University of Technology53 (DU-W family). More
recently, Fuglsang and Dahl54 at Risù used optimization techniques for special-purpose aerofoil design.
With respect to the use of thick aerofoils, it has become clear that upscaling in thickness of the general
aviation aerofoils gives very poor performance. Special pro®les of relative thickness of up to 30% have
been designed at both FFA and DUT and are presently used by the main blade manufacturers. An
inherent property of these thick aerofoils is the small a range between the design value and the Cl,max
value. On the other hand, thinner pro®les have a much wider a range in this context, especially the general
aviation families. Since, moreover, the change in a with wind speed (for a ®xed speed, ®xed pitch turbine)
is much larger in the root than at the tip, it follows that a thick root section will stall too soon and a thin
tip section will stall too late. This is resolved by some manufacturers by using vortex generators in the root
section (up to 50% span) and using stall strips in the tip section. Another approach is to design thin
aerofoils speci®cally for a smaller a range. This also is one item of research and development.
In general, since blade design details are of a competitive nature, not much information is present in the
open literature with regard to these items. However, it is clear that the application of vortex generators or
other boundary layer manipulators on rotating blades is done in an empirical manner and that much
could be learned by systematic (wind tunnel) investigations.

Acknowledgements
The author wishes to express his gratitude to many of his colleagues who have given information and
comments that have improved the quality of this review. In particular, thanks go to Gustave Corten and
Gerard Schepers.

Appendix: System of Flow Equations


Introduction
The derivation of the equations describing the ¯ow of ¯uids can be found in many textbooks and
monographs, e.g. Reference 55. In this appendix the equations are stated and interpreted physically to aid
in the understanding of the basis of certain approximations that are discussed in the main text.
Space co-ordinates will be given in a Cartesian system xj , j ˆ 1, 2, 3, and the velocity ®eld will be given
by the three velocity components uj , which are functions of position xj and time t. Other important ¯ow
quantities are the pressure p and the mass density r.
Throughout this appendix, incompressible ¯ow of uniform density will be assumed. The approximation
of incompressibility is valid for ¯ow ®elds in which the local velocities are small compared with the speed
of sound propagation (small Mach number). In that case, density variations resulting from pressure
variations can be neglected. The speed of sound a in an ideal gas (e.g. air at atmospheric conditions) is
given by
p cp
a ˆ gRT ; g ˆ ; R ˆ cp ÿ cv
cv

# 1998 John Wiley & Sons, Ltd. Wind Energ., 1, 46±69 (1998)
62 H. Snel

where cp and cv are the gas (air) speci®c heat coecients for contrast pressure and constant volume
processes respectively, R is the gas constant and T denotes absolute temperature (K). For air, g ˆ 1.4 and
R ˆ 287 J kg71 K71. Hence for normal atmospheric temperatures a will be equal to approximately
340 m s71. For local ¯ow speeds up to 100 m s71 the approximation of incompressibility is acceptable.
For the density to be uniform, the additional requirement exists that temperature and strati®cation e€ects
on the density must also be neglected.

Navier±Stokes Equations
Under the conditions outlined above, the Navier±Stokes equations describing the ¯ow of a ¯uid are

@uj
ˆ0 …9†
@xj

@ui @u 1 @p @2 u
‡ uj i ˆ ÿ ‡ v 2i ‡ f i ; i ˆ 1; 2; 3 …10†
@t @xj r @xi @xj

The usual convention is used that a repeated index in a single term implies summation over the index
values.
The scalar equation (9) is known as the continuity equation, expressing mass conservation. For the
constant density case this is equivalent to volume conservation. Hence the net volume ¯ow across a closed
surface must be zero or mathematically expressed: the velocity ®eld is divergence-free. This is expressed by
equation (9).
The vector equation (10) is Newton's law of conservation of momentum in the three co-ordinate
directions. This is referred to as the momentum equation. The two terms on the left-hand side (IHS) of the
equation express the acceleration of a ¯uid particle. The partial time derivative is known as the local
acceleration and the second, non-linear term describes the acceleration due to convection of the particle
with the ¯ow ®eld (convective acceleration). The right-hand side (RHS) of the equation contains the
forces (per unit mass) that are responsible for the accelerations. The ®rst term is the force due to pressure
di€erences and the second term represents the viscous force. The quantity v is the kinematic viscosity of
the ¯uid. Its value is a slowly varying function of the temperature; for `normal' atmospheric air its value is
approximately equal to 1.5  1075 m2 s71. Finally, the term fi represents a possible external (body) force
per unit mass.
The solutions to these equations must satisfy certain boundary conditions. On solid surfaces within the
¯ow ®eld the ¯ow velocity relative to that surface must be equal to zero. This can be conveniently
decomposed into a `no-transparency' condition (normal velocity zero) and a `no-slip' condition
(tangential velocity zero). It should be noted that these conditions are not prescribed on an actuator disc.
Instead, external forces applied at such a surface are to be prescribed, but the ¯ow can pass through and
by an actuator disc.
The total system consists of four scalar equations in four unknowns, namely the velocity vector
components ui and the (scalar) pressure p. It is non-linear in the velocity components through the
convective acceleration terms. Analytical solutions are known for very few special cases. Direct numerical
solution of the equations for ¯ows with large Reynolds numbers (see below) has been impossible for one
main reason: the ¯ow develops instabilities in regions of large shear and ¯uctuations with time and
position occur on such small scales that no computer power is able to handle these.

# 1998 John Wiley & Sons, Ltd. Wind Energ., 1, 46±69 (1998)
Review of Rotor Aerodynamics 63

The Vorticity Transport Equation


An alternative form of the momentum equation can be obtained by casting it in terms of the vorticity
vector ok , which is de®ned mathematically as the curl of the velocity vector ui :
@
ok ˆ  ui …11†
@xj

The vorticity is directly related to the rotational velocity of ¯uid particles, the rotational axis being along
the direction of the vorticity vector.
Taking the curl of equation (10) and making use of (9), an equation is obtained which is usually called
the vorticity transport equation:

@oi @o @u @2 o @
‡ uj i ˆ oj i ‡ v 2i ‡  fk …12†
@t @xj @xj @xj @xj

The IHS of the equation only describes the rate of change in vorticity following a material particle; the
RHS describes the causes of the change. The ®rst term on the RHS is the most dicult to describe
physically. It expresses the redistribution of vorticity by deformation of vortex lines, such as stretching
and rotation. Note that this term is equal to zero in the 2D case, since then the vorticity vector is normal
to the plane of ¯ow, while the velocity derivative in that normal direction equals zero. The second term
describes the viscous di€usion of vorticity, as is clear from its form, which is that of a Laplace operator. It
is similar to a conduction term if the quantity is a scalar, e.g. temperature. It should be kept in mind,
however, that vorticity is a vector quantity and a function of the velocity ®eld ui , so equation (12) is
strongly non-linear. The last term describes the creation of vorticity by the body forces.
In fact, vorticity can be created at solid boundaries as a result of the no-slip condition or through the
action of external forces. With respect to these external forces the reservation should be made that they
can only create vorticity if they are not derived from a force potential (i.e. if they are non-conservative),
since in that case the curl would be equal to zero. This in fact made the pressure term disappear on taking
the curl of (10). Physically, this becomes clear by observing that pressure forces are normal (to the particle
surface) and cannot cause rotation. The external forces at an actuator disc will usually create vorticity.

The Reynolds Number


As a result of the small value of viscosity, the viscous forces expressed in equation (10) will only be
important in regions where the spatial derivatives of the velocity (velocity gradients) are large, or more
precisely, where the variations in these gradients are large. Equivalently, the di€usion of vorticity will
only be important in regions of high vorticity gradients. This can be expressed conveniently through the
use of the Reynolds number, as will be discussed here.
In fact, let U be a representative value of the velocity and let L be a representative length scale of the
problem. If spatial derivatives are of the order of U/L, then the order-of-magnitude relation between the
convective acceleration and the viscous forces (per unit mass) can be inferred from (10) to be
" #
@ui
O uj U2
@xj UL
" #ˆ L ˆ ˆ Re …13†
2
@ ui U v
O v 2 v
@xj L2

The Reynolds number Re is an important non-dimensional quantity that characterizes the type of ¯ow.
For the ¯ow about a wind turbine the characteristic velocity is of order Or and for the length scale it is

# 1998 John Wiley & Sons, Ltd. Wind Energ., 1, 46±69 (1998)
64 H. Snel

customary to use the aerofoil chord length c. A typical value of the Reynolds number for a large turbine
(e.g. c ˆ 1.5 m, Or ˆ 50 m s71) will be around 5  106. This means that the accelerations are very large in
comparison with the viscous forces, so the latter cannot be responsible for the accelerations. Hence the
pressure forces must be dominating the ¯ow ®eld. This considerations leads to the approximation of ¯ow
without viscosity.

The Euler Equations


Neglecting the viscous terms in equation (10), the so-called Euler equations are obtained as
@ui @u 1 @p
‡ uj i ˆ ÿ ‡ f i; i ˆ 1; 2; 3 …14†
@t @xj r @xi

together with the unchanged equation (9).


These equations are of lower order in the spatial derivatives than the Navier±Stokes equations and
cannot be made to obey the same boundary conditions. In fact, solutions to the Euler equations are only
supposed to satisfy the `no-transparency' boundary condition. This also implies that no vorticity is
generated in the ¯ow ®eld, apart from the possible action of the external body forces. However, this is in
disagreement with physical reality, as will be discussed below.
If the ¯ow is emanating from a vorticity-free region (e.g. a region of uniform velocity), then the entire
¯ow will be vorticity-free (or irrotational) under the assumptions that the Euler equations are a correct
model and that no vorticity is created by body forces. In this case the velocity ®eld can be described even
more simply by noting that a vector quantity whose curl equals zero can be expressed as the gradient of a
scalar function. Hence
@
ui ˆ F …15†
@xi

where F is known as the velocity potential. Since the divergence of ui equals zero, F satis®es the Laplace
equation

@2 F
ˆ0 …16†
@x2i

which has to satisfy the non-transparency condition on solid surfaces, i.e.


@F=@n ˆ un ˆ 0 …17†

where n is the unit direction vector normal to the surface. This ¯ow model is known as potential ¯ow
(inviscid, vorticity-free and incompressible). It is possible to resolve the velocity ®eld by solving the
Neumann problem for the Laplace equation. The pressure can then be obtained from the momentum
equation (14).
When applying the Euler equations to the global ¯ow ®eld about a turbine rotor, the ¯ow upstream of
the rotor may be regarded as vorticity-free. However, vorticity is created at the actuator disc which models
the rotor, through the external forces prescribed there. In the wake, downstream of the rotor, this vorticity
is transported, but the Euler equation does not model its di€usion. Because of the large Reynolds number
involved and the absence of solid surfaces, this is a valid approximation in general.
However, when applying inviscid ¯ow theory to the ¯ow about the solid blade surface, the following
problem arises. Vorticity is formed physically on the actual blade surface as a result of the no-slip
condition, but this condition is not applicable to the inviscid equations (14) and (16). In reality a thin
region on the boundary will exist where the velocity gradients are so high that the viscous terms are of the
same order of magnitude as the pressure terms and the acceleration. This region, known as the boundary

# 1998 John Wiley & Sons, Ltd. Wind Energ., 1, 46±69 (1998)
Review of Rotor Aerodynamics 65

layer, contains all the vorticity and is extended into the wake. It is thin because of the fact that the
convectional velocity of the vorticity in the direction parallel to the surface is very much higher than the
`di€usion velocity' in the direction normal to the surface. A ®x to the problem of vorticity formation and
di€usion in high-Reynolds-number ¯ows is found in Prandtl's boundary layer theory.

Boundary Layer Equation


In the boundary layer approximation the ¯ow ®eld is modelled in two parts. One is the boundary layer
directly adjacent to a solid boundary and its ensuing wake, containing vorticity. The other part is usually
called the `outer ¯ow', which is modelled as vorticity-free. Within the boundary layer, let x2 denote the
direction normal to the solid surface and x1 and x3 be directions along the surface; see Figure 4 (surface
curvature can be neglected usually).

Figure 4. Co-ordinate directions used for boundary layer description

In the boundary layer approximation the pressure change across the boundary layer is neglected, i.e. the
momentum equation for direction x2 reduces to
@p
ˆ0 …18†
@x2

For the momentum equations parallel to the surface the viscous terms containing derivatives in directions
x1 and x3 are neglected (in these directions, di€usion is very small compared with convection) and only
the second derivative in the direction normal to the surface is retained, since in that direction the velocity
gradient is very large indeed. Hence these equations read

@ui @u 1 @p @2 u
‡ uj i ˆ ÿ ‡ v 2i ; i ˆ 1; 3; j ˆ 1; 2; 3 …19†
@t @xj r @xi @x2

Note that the convective terms of the LHS of the equation remain completely 3D in character.
The boundary layer equations are of parabolic type, as opposed to the elliptic character of the Navier±
Stokes equations. This means that the boundary layer equations can be solved by a marching procedure in
the streamwise direction, starting from `initial' conditions and applying boundary conditions on the body
(zero relative velocity) and on the interface between the body and the outer ¯ow (velocity equality). There
is no ¯ow of information in the direction opposed to the velocity direction. The Navier±Stokes equations,
however, need boundary conditions all around the ¯ow ®eld, and information can travel in all directions.
This results in far more ecient numerical solutions schemes for the boundary layer equations in
comparison with the Navier±Stokes equations.
The boundary layer equations are solved with known pressure gradient terms along the surface. These
latter are obtained from an inviscid solution of the outer ¯ow. In this solution, however, the boundary
layer vorticity must be present by introducing a quantity of vorticity, which is placed along the boundary
or within a surface surrounded by the boundary. The outer ¯ow ®eld is assumed to be vorticity-free. In
fact, the inviscid ¯ow solution can be seen as the limit of the Navier±Stokes solution for Re approaching

# 1998 John Wiley & Sons, Ltd. Wind Energ., 1, 46±69 (1998)
66 H. Snel

in®nity, in which the thickness of the boundary layer approaches zero and the velocity gradient in the
same approaches a discontinuity, from zero at y ˆ 0 to the `potential ¯ow' value for y ˆ 0 ‡ .
A problem arises when the boundary layer separates from the surface and forms a wake of signi®cant
size. The boundary layer equations with prescribed pressure terms become singular at separation. A
di€erent solution strategy must be followed, resolving the boundary equations simultaneously with the
outer ¯ow equations, accounting for the displacement e€ects on the outer ¯ow. This is known as strong
interaction. It enables the construction of solutions for ¯ows with small separation regions. However,
when separation becomes massive, the region containing vorticity is no longer thin compared with the
main ¯ow dimensions (i.e. an aerofoil chord), and the assumptions underlying the boundary layer theory
are no longer valid.

Turbulence and Reynolds Averaging


Turbulence results from intrinsic unstable behaviour of the ¯ow that develops in regions with large shear
at suciently large Reynolds numbers, when viscosity is no longer able to provide sucient damping of
¯ow perturbations. Under these circumstances the ¯ow becomes time-dependent even under steady state
external conditions. For unsteady external conditions it is possible to make a distinction between the
small-scale turbulent ¯uctuations and those that arise from the external conditions. Then a decom-
position is made into the `average ¯ow' and the turbulent ¯uctuations:
0
ui ˆ u i ‡ ui …20†

where the `barred' quantity is the so-called Reynolds-averaged quantity and the `primed' quantity is the
turbulent ¯uctuation. In theory the `Reynolds average' should be an ensemble average, i.e. the average
over a large (in®nite) number of realizations with the same external conditions. Usually one wants a
solution for the average ¯ow ®eld, but this is in¯uenced by the presence of the turbulent ¯uctuations.
This interaction becomes clear by using the Reynolds-averaged Navier±Stokes (RANS) equations. These
are obtained by substituting the decomposed velocities and pressures into equations (9) and (10) and
subsequently Reynolds averaging the result. Here only the result will be shown for the momentum
equations (the continuity equation is unchanged):
 
@ u 0 u0
@u i @ 
u j 1 @p 2
@ u i j
‡ u j ˆÿ ‡ v 2i ÿ …21†
@t @xj r @xi @xj @xj

The last term in this equation results from the non-linear convective acceleration term of (10), after
applying (9) and a Reynolds average. It is known as the Reynolds stress or turbulent shear stress term and
represents i-momentum exchange through the action of the turbulent velocity ¯uctuations. It has the
form of a second-order correlation. It is usually very much larger than the viscous shear term. In e€ect
the viscous term can be interpreted as momentum exchange through molecular speed ¯uctuations, while
the turbulent term is the result of much larger-scale correlated motions. Usually the viscous term is
neglected except in a region extremely close to a solid wall, divided into the viscous sublayer and the
bu€er layer. In the ®rst layer the presence of the wall inhibits the turbulence to such a degree that
viscosity is dominant; the second layer is de®ned as the region where the two mechanisms of momentum
exchange are of equal order of magnitude.
The Reynolds stress term represents a new unknown in the equations for the average ¯ow ®eld. It is
possible to derive equations for the Reynolds stresses by basically the same procedure (after multiplying
the i-momentum equation by uj), but these equations will include new unknowns in the form of third-
order correlations. This is known as the closure problem. The only way to get a closed system of Reynolds-
averaged equations is by assuming other types of relations between the Reynolds stresses (or higher-order
correlations) and the average ¯ow ®eld. These relations are known as turbulence models. Many have been

# 1998 John Wiley & Sons, Ltd. Wind Energ., 1, 46±69 (1998)
Review of Rotor Aerodynamics 67

derived (on empirical grounds), but no conclusive evidence exists for the practical correctness of any of
these for a wide variety of ¯ow conditions. In particular, separating high-Reynolds-number boundary
layers present a problem.

Notation

a axial induction factor (V 7 VD)V ˆ ui/V


AR swept rotor area or, equivalently, actuator disc area, pD2/4 (m2)
c aerofoil chord (m)
Cl lift coecient, (aerofoil lift) 12rV 2 c
Cd drag coecient, (aerofoil drag) 12rV 2 c
CD,ax axial force coecient, Dax =AR 12rV 2
CP power coecient, P=AR 12rV 3
CP,max maximum value of CP at design conditions
D rotor diameter (m)
Dax axial force on rotor (N)
r rotor radius, D/2 (m)
ui induced velocity in rotor plane (m s71)
V undisturbed wind speed (m s71)
VD wind speed at rotor disc position (m s71)
Vef local e€ective velocity (m s71)
Vw wind speed in far wake (m s71)

Greek Letters
a angle of attack (8)
G vortex strength (m2 s71)
l tip speed ratio, Or/V
r air density, approximately equal to 1.25 at sea level (kg m73)
O rotor rotational speed (rad s71)

References
1. A. C. Hansen and C. P. Butter®eld, `Aerodynamics of horizontal-axis wind turbines', Ann. Rev. Fluid Mech., 25,
115±149 (1993).
2. H. Snel and Th. van Holten, `Review of recent aerodynamical research on wind turbines with relevance to
rotorcraft', Aerodynamics and Aerocoustics of Rotorcraft, AGARD CP 552, August 1995, pp. 7±11.
3. J. N. Sùrensen, `A survey of CFD methods in rotor aerodynamics', 7th IEA Symp. on Aerodynamics of Wind
Turbines, Lyngby, November 1993.
4. R. E. Froude, `On the part played in propulsion by di€erences of ¯uid pressure', Trans. Inst. Naval Archit., 30
(1889).
5. F. W. Lanchester, `A contribution to the theory of propulsion and the screw propeller', Trans. Inst. Naval
Archit., 57 (1915).
6. A. Betz, `Das Maximum der theoretisch moÈglichen AusnuÈtzung des Windes durch Windmotoren', Z. Gesamte
Turbinewesen, 26 (1920).
7. G. A. M. van Kuik, `On the limitations of Froude's actuator disc', Doctoral Thesis, University of Eindhoven,
1991.
8. O. Igra, `Research and development for shrouded wind turbines', Proc. EWEC, Hamburg, 1984.
9. Th. van Holten, `Windmills with di€user e€ect induced by small tip vanes', Proc. Int. Symp. on Wind Energy
Systems, Cambridge, 1976.
10. O. de Vries, `Fluid dynamic aspects of wind energy conversion', AGARDograph AG-243, 1979.

# 1998 John Wiley & Sons, Ltd. Wind Energ., 1, 46±69 (1998)
68 H. Snel

11. H. Glauert, `Windmills and fans', in W. F. Durand (ed.), Aerodynamic Theory, Vo.. 14, Division L, Dover,
New York, 1963, pp. 324±340.
12. R. E. Wilson and P. B. S. Lissaman, `Applied aerodynamics of wind power machines', Oregon State University
Report NSF/RA/N-74113, 1974.
13. P. Goorian, `An invalid equation in the general momentum theory of the actuator disk', AIAA, 10(4), 543±544,
April 1972.
14. A. C. Hansen, `Yaw dynamics of horizontal axis wind turbines: ®nal report', NREL Technical Report, 442±4822,
19.
15. J. G. Schepers and H. Snel, `Dynamic in¯ow: yawed conditions and partial span pitch control', ECN-C-95-056,
1995.
16. G. H. Goankar and D. A. Peters, `Review of dynamic in¯ow modelling for rotorcraft ¯ight mechanics', Vertica,
12(3), 213±242 (1988).
17. H. Snel and J. G. Schepers, `Joint investigation of dynamic in¯ow e€ects and implementation of an engineering
method', ECN-C-94-107, 1994.
18. H. J. van Grol, H. Snel and J. G. Schepers, `Wind turbine benchmark exercise on mechanical loads, Volume 1,
Part A and B. A state of the art report', ECN-C-91-031, 1991.
19. H. A. Madsen, `Application of actuator surface theory on wind turbines', Proc. 2nd IEA Symp. on Aerodynamics
of Wind Turbines, Lyngby, 1988.
20. F. J. Simoes and J. M. R. Graham, `Prediction of loading on a horizontal axis wind turbine using a free vortex
wake model', Proc. BWEA Conf., 1991.
21. R. Bareiû, G. Guidati and S. Wagner, `Wake simulation for wind turbines with a free, prescribed and hybrid
wake method', Proc. 10th IEA Symp. on Aerodynamics of Wind Turbines, Edinburgh, 1996 (accepted for
publication).
22. S. G. Voutsinas, `Development of a new generation of design tools for HAWT', Final Report of the JOU2-CT92-
0113 CEC Project, 1995.
23. H. Snel, `Some observations concerning blade-element-momentum (BEM) methods and vortex wake methods,
including numerical experiments', Proc. 10th IEA Symp. on Aerodynamics of Wind Turbines, Edinburgh, 1996,
pp. 3±22.
24. N. J. Vermeer, `How big is a tip vortex?', Proc. 10th IEA Symp. on Aerodynamics of Wind Turbines, Edinburgh,
1996, pp. 77±82.
25. Th. van Holten, `On the validity of lifting line concepts in rotor analysis', Vertica, 1, (1977).
26. G. J. W. van Bussel, `The aerodynamics of horizontal axis wind turbines rotors explored with asymptotic
expansion methods'. Doctoral Thesis, Delft University, 1995.
27. G. J. W. van Bussel, `The application of advanced rotor (performance) methods for design calculations', Proc.
10th IEA Symp. on Aerodynamics of Wind Turbines, Edinburgh, 1996 (accepted for publication).
28. H. A. Madsen, `A CFD analysis of the actuator disc ¯ow compared with momentum theory results', Proc. 10th
IEA Symp. on Aerodynamics of Wind Turbines, Edinburgh, 1996 (accepted for publication).
29. A. BjoÈrck, `Dynamic stall and three-dimensional e€ects', Final Report for the EC DGXII Joule II Project JOU2-
CT93-0345, FFA TN 1995-31, 1995.
30. V. K. Truong, `A 2D dynamic stall model based on a Hopf bifurcation', Nineteenth European Rotorcraft Forum,
Cernobbio, 1993, Paper C23.
31. J. Gould and S. P. Fiddes, `Computational methods for the performance prediction of HAWTs', Proc. EWEC,
Amsterdam, 1991, pp. 29±33.
32. H. A. Madsen, C. Bak, P. Fuglsang and F. Rasmussen, `The phenomenon of double stall', Proc. EWEC, Dublin,
1997.
33. J. L. van Ingen, `A suggested semi-empirical method for the calculation of the boundary layer region', Report
VTH-74, Delft University of Technology, Dept. of Aeronautical Engineering, 1956.
34. A. M. O. Smith, `Transition, pressure gradient and stability theory', Proc. IX Int. Congr. of Applied Mechanics,
Brussels, 1956.
35. W. P. Jones and B. E. Launder, `The calculation of low Reynolds number phenomena with a two equation model
of turbulence', Int. J. Heat Mass Transfer, 16, 1119±1130 (1973).
36. D. C. Wilcox, Turbulence Modeling for CFD, DCW Industries, La Canada, CA, 1993.
37. F. R. Menter, `Zonal two equation k±o models for aerodynamic ¯ows', AIAA Paper 93-2906, 1993.
38. G. R. Srinivasan, J. A. Ekaterinaris and W. J. McCroskey, `Evaluation of turbulence models for unsteady ¯ows
on an oscillating airfoil', Comput. Fluids, 24, (1995).
39. J. N. Sùrensen, `Three level viscous±inviscid interaction technique for the prediction of separated ¯ow past
rotating wing', PhD Thesis, AFM-83-03, Technical University of Denmark, 1986.
40. P. Bunniss, `Detailed measurements in the boundary layer by LDA. Experimental data', University of Bristol,
Department of Aerospace Engineering Report 737, 1995.

# 1998 John Wiley & Sons, Ltd. Wind Energ., 1, 46±69 (1998)
Review of Rotor Aerodynamics 69

41. H. Snel, `Scaling laws for the boundary layer ¯ows on rotating wind turbine blades', Proc. 4th IEA Symp. on
Aerodynamics of Wind Turbines, Rome, 1991.
42. H. Snel, R. Houwink and J. Bosschers, `Sectional prediction of lift coecients on rotating wind turbine blades in
stall', ECN-C-93-052, 1994.
43. W. P. Wolfe and S. S. Ochs, `Predicting aerodynamic characteristics of typical wind turbine airfoils using CFD',
Sandia Report SAND96-2345, 1997.
44. G. M. Gregorek, M. J. Ho€man and M. J. Berchak, `Steady state and oscillatory aerodynamic characteristics of
a NACA 0012 airfoil', Data Report, Ohio State University, 1989.
45. P. K. Chaviaropoulos, `Investigating dynamic stall, 3-D and rotational e€ects on wind turbine blades by means
of an unsteady quasi-3D Navier Stokes solver', Proc. 10th IEA Symp. on Aerodynamics of Wind Turbines,
Edinburgh, 1996, pp. 175±182.
46. M. O. L. Hansen, N. N. Sùrensen, J. N. Sùrensen and J. A. Michelsen, `A global Navier Stokes rotor prediction
model', AIAA paper 97-0970, 1997.
47. R. L. Hales, `Dynamic stall on horizontal axis wind turbines', Proc. EWEC, Amsterdam, 1991, pp. 34±39.
48. J. G. Schepers, A. J. Brand, A. Bruining, J. M. R. Graham, M. M. Hand, D. G. In®eld, H. A. Madsen, R. J. H.
Paynter and D. A. Simms, Final Report of IEA Annex XIV: Field Rotor Aerodynamics, ECN-C-97-027, 1997.
49. G. Ronsten, `Static pressure measurements on a rotating and non rotating 2.375 m wind turbine blade.
Comparison with 2D calculations', J. Wind Engng. Ind. Aerodyn., 39, (1992).
50. M. J. Ho€man, R. R. Ramsay and G. M. Gregorek, `Unsteady aerodynamic performance of wind turbine
airfoils', Proc. AWEA Wind Power 94 Conf., Minneapolis, MN, May 1994, pp. 583±594.
51. J. L. Tangler and D. M. Somers, `Status of the special purpose airfoil families', Windpower '87 Conf., San
Francisco, CA, 1987, pp. 99±105.
52. A. BjoÈrck, `Co-ordinates and calculations for the FFA-WI-xxx, FFA-W2-xxx and FA-W3-xxx series of airfoils
for horizontal axis wind turbines', FFA TN 1990-15, 1990.
53. W. A. Timmer and R. P. J. O. M., van Rooij, `The performance of new wind turbine blade tip and root airfoils
up to high angles of attack', Proc. EWEC, Dublin 1997 (accepted for publication).
54. P. Fuglsang and K. S. Dahl, `Multipoint optimisation of thick high lift airfoil for wind turbines', Proc. EWEC,
Dublin, 1997.
55. F. M. White, Viscous Flow Theory, McGraw-Hill, New York, 1974.

# 1998 John Wiley & Sons, Ltd. Wind Energ., 1, 46±69 (1998)

You might also like