You are on page 1of 17

2.

08 Aerodynamic Analysis of Wind Turbines


JN Sørensen, Technical University of Denmark, Lyngby, Denmark
© 2012 Elsevier Ltd. All rights reserved.

2.08.1 Introduction 225

2.08.2 Momentum Theory 226

2.08.2.1 One-Dimensional Momentum Theory 226

2.08.2.2 The Optimum Rotor of Glauert 227

2.08.2.3 The Blade-Element Momentum Theory 228

2.08.2.3.1 Tip correction 229

2.08.2.3.2 Correction for heavily loaded rotors 230

2.08.2.3.3 Yaw correction 230

2.08.2.3.4 Dynamic wake 230

2.08.2.3.5 Airfoil data 231

2.08.3 Advanced Aerodynamic Modeling 231

2.08.3.1 Vortex Models 231

2.08.3.2 Numerical Actuator Disk Models 232

2.08.3.3 Full Navier–Stokes Modeling 232

2.08.4 CFD Computations of Wind Turbine Rotors 233

2.08.5 CFD in Wake Computations 234

2.08.6 Rotor Optimization Using BEM Technique 236

2.08.7 Noise from Wind Turbines 238

References 239

Further Reading 240

2.08.1 Introduction

The aerodynamics of wind turbines concerns, briefly speaking, modeling and prediction of the aerodynamic forces on the solid
structures of a wind turbine and in particular on the rotor blades of the turbine. Aerodynamics is the most central discipline for
predicting performance and loadings on wind turbines. The aerodynamic model is normally integrated with models for wind
conditions and structural dynamics. The integrated aeroelastic model for predicting performance and structural deflections is a
prerequisite for design, development, and optimization of wind turbines. Aerodynamic modeling may also concern design of
specific parts of wind turbines, such as rotor blade geometry or performance predictions of wind farms.
Using simple axial momentum theory and energy conservation, Lanchester [1] and Betz [2] predicted that even an ideal wind
turbine cannot exploit more than 59.3% of the wind power passing through the rotor disk. A major breakthrough in rotor
aerodynamics was achieved by Betz [2] and Glauert [3], who formulated the blade-element momentum (BEM) theory. This theory,
which later has been extended with many ‘engineering rules’, is today the basis for all rotor design codes in use by industry.
From an outsider’s point of view, aerodynamics of wind turbines may seem simple as compared to aerodynamics of, for
example, fixed-wing aircraft or helicopters. However, there are several added complexities. Most prominently, aerodynamic stall is
always avoided for aircraft, whereas it is an intrinsic part of the wind turbines operational envelope. Stall occurs when the flow meets
the wing at a too high angle of attack. The flow then cannot follow the wing surface and separates from the surface, leading to flow
patterns far more complex than that of nonseparated flow. This renders an adequate description very complicated, and even for
Navier–Stokes simulations, it becomes necessary to model the turbulent small-scale structures in the flow, using Reynolds-averaging
or large eddy simulations (LESs). Indeed, in spite of the wind turbine being one of the oldest devices for exploiting the energy of the
wind, some of the most basic aerodynamic mechanisms are not yet fully understood.
Wind turbines are subjected to atmospheric turbulence, wind shear from the ground effect, wind directions that change both in
time and in space, and effects from the wake of neighboring wind turbines. These effects together form the ordinary operating
conditions experienced by the blades. As a consequence, the forces vary in time and space and a dynamical description is an intrinsic
part of the aerodynamic analysis.
At high wind velocities, where a large part of the blade of stall-regulated turbines operates in deep stall, the power output is
extremely difficult to determine within an acceptable accuracy. When boundary layer separation occurs, the centrifugal force tends
to push the airflow at the blade toward the tip, resulting in the aerodynamic lift being higher than what it would be on a nonrotating
blade.
When the wind changes direction, misalignment with the rotational axis occurs, resulting in yaw error. Yaw error causes periodic
variation in the angle of attack and invalidates the assumption of axisymmetric inflow conditions. Furthermore, it gives rise to radial
flow components in the boundary layer. Thus, both the airfoil characteristics and the wake are subject to complicated
three-dimensional (3D) and unsteady flow behavior.

Comprehensive Renewable Energy, Volume 2 doi:10.1016/B978-0-08-087872-0.00209-2 225


226 Aerodynamic Analysis of Wind Turbines

In the following, a brief introduction is given to wind turbine aerodynamics. It is not possible in a short form to introduce to all
aspects of rotor aerodynamics and the scope is on conventional aerodynamic modeling, as it is still used by industry in the design of
new turbines, and on state-of-the-art methods for analyzing wind turbine rotors and wakes. Specifically, the basics of momentum
theory, which still form the backbone in rotor design of wind turbines, are introduced. Next, state-of-the-art advanced aerodynamic
models is presented. This includes vortex models, generalized actuator disk/line models, and computational fluid dynamics (CFD).
Finally, a short introduction is given to rotor optimization and modeling of aerodynamically generated noise.

2.08.2 Momentum Theory

The basic tool for understanding wind turbine aerodynamics is the momentum theory in which the flow is assumed to be inviscid,
incompressible, and axisymmetric. The momentum theory consists basically of control volume integrals for conservation of mass,
axial and angular momentum balances, and energy conservation. In the following, we will give a brief introduction to momentum
theory for design and analysis of wind turbines, starting by the simple, albeit important, one-dimensional (1D) momentum theory,
from which the Betz limit can be derived, and ending with the practical BEM theory, which forms the basis for all rotor design codes
in use by industry.

2.08.2.1 One-Dimensional Momentum Theory


We first revisit the simple axial momentum theory as originated by Rankine [4], Froude [5], and Froude [6]. Consider an axial flow
of speed Uo passes through an actuator disk of area A with constant axial load (thrust) T. Denoting by uR the axial velocity in the
rotor plane, and let u1 be the axial velocity in the ultimate wake where the air has regained its undisturbed pressure value, pw = po,
and let ρ denote the density of air. We now consider a 1D model for the stream tube that encloses the rotor disk (see Figure 1), and
denote by Ao and A1 the cross-sectional area of the flow far upstream and far downstream of the rotor, respectively.
The equation of continuity requires that the rate of mass flow, ṁ, is constant in each cross-section. Thus,

ṁ ¼ ρUo Ao ¼ ρuR A ¼ ρu1 A1 ½1


Axial momentum balance for the considered stream tube results in the following equation for the thrust

T ¼ ṁðUo − u1 Þ ¼ ρuR AðUo − u1 Þ ½2


Applying the Bernoulli equation in front of and behind the rotor, we find that the total pressure head of the air in the slipstream has
been decreased by
1  
Δp ¼ ρ Uo2 − u21 ½3
2
The pressure drop takes place across the rotor and represents the thrust, T = AΔp. Combining eqns [2] and [3] shows the well-known
result that
1
uR ¼ ðu1 þ Uo Þ ½4
2
Introducing the axial interference factor as follows:

Uo − uR
a¼ ½5
Uo

we obtain uR = (1 − a)Uo and u1 = (1 − 2a)Uo. From eqn [2], we get the following expressions for thrust and power extraction:

Uo uR u1

Figure 1 Control volume for 1D actuator disk.


Aerodynamic Analysis of Wind Turbines 227

T ¼ 2ρAUo2 að1− aÞ ½6

P ¼ uR T ¼ 2ρAUo2 að1− aÞ 2 ½7


Introducing the dimensionless thrust and power coefficient, respectively,

T P
CT ≡ ; CP ≡ ½8
2 ρAUo 2 ρAUo
1 2 1 3

we get
CT ¼ 4að1− aÞ; CP ¼ 4a ð1 − a Þ 2 ½9
Differentiating the power coefficient with respect to the axial interference factor, the maximum obtainable power is obtained as
16 1
CPmax ¼ ¼ 0:593 for a¼ ½10
27 3
This result is usually referred to as the Betz limit or the ‘Lanchester–Betz–Joukowsky limit’, as recently proposed by van Kuik [7], and
states the upper maximum for power extraction which is no more than 59.3% of the kinetic energy contained in a stream tube
having the same cross-section as the disk area can be converted to useful work by the disk. However, it does not include the losses
due to rotation of the wake and therefore it represents a conservative upper maximum.

2.08.2.2 The Optimum Rotor of Glauert


Utilizing general momentum theory, Glauert developed a simple model for the optimum rotor that included rotational velocities.
In this approach, Glauert treated the rotor as a rotating axisymmetric actuator disk, corresponding to a rotor with an infinite number
of blades. The main approximation in Glauert’s analysis was to ignore the influence of the azimuthal velocity and pressure in the
axial momentum equation. For a differential element of radial size Δr, eqn [2] then reads,

ΔT ¼ 4πρUo2 ð1− aÞarΔr ½11


Applying the Bernoulli equation in a rotating frame of reference across the rotor plane, we get the following equation for the
pressure drop over the rotor,
1
Δp ¼ ρΩruθ þ ρu2θ ½12
2
where Ω is the angular velocity.
Combining eqns [11] and [12], we get
 
1
ΔT ¼ Δp ⋅ 2πrΔr ¼ 2πρuθ Ωr þ uθ rΔr ½13
2
where uθ is the azimuthal velocity behind the rotor. Defining the azimuthal interference factor as,

a′¼ ½14
2Ωr
eqn [13] reads,

ΔT ¼ 4πρΩ2 ð1 þ a′Þa′r 3 Δr ½15


Combining eqns [11] and [15], we get
ð1− aÞa ¼ λ2 x2 ð1 þ a′Þa′ ½16
where x = r/R and λ = ΩR/Uo is the tip speed ratio. This equation can also be derived by letting the induced velocity be perpendicular
to the relative velocity in the rotor plane. Introducing Euler’s turbine equation on differential form, we get the following expression
for the useful power produced by the wind turbine,

ZR Z1
P¼Ω 2πr ρuuθ dr ¼ 4πρΩ Uo
2 2
a′ð1− aÞx3 dx ½17
0 0

or in dimensionless form,

Z1
CP ¼ 8λ2 a′ð1− aÞx3 dx ½18
0
228 Aerodynamic Analysis of Wind Turbines

By assuming that the different stream tube elements behave independently of each other, it is possible to optimize the integrand for
each x separately (see Glauert [3] or Wilson and Lissamann [8]). This results in the following relation for an optimum rotor,

da′
ð1− aÞ − a′ ¼ 0 ½19
da
Differentiating eqn [16] with respect to a gives,

da′
1− 2a ¼ λ2 x2 ð1 þ 2a′Þ ½20
da
Combining eqns [16], [19], and [20] results in the following relationship

1 − 3a
a′ ¼ ½21
4a − 1
The analysis shows that the optimum axial interference factor is no longer a constant but will depend on the rotation of the wake
and that the operating range for an optimum rotor is 1/4 ≤ a ≤ 1/3.
The relations between a, a′, a′x2λ2, and λx for an optimum rotor are given in Table 1. The maximal power coefficient as a function
of tip speed ratio is determined by integrating eqn [18] and is shown in Table 2. The optimal power coefficient approaches 0.593 at
large tip speed ratios only. It shall be mentioned that these results are valid only for a rotor with an infinite number of blades and
that the analysis is based on the assumption that the rotor can be optimized by considering each blade element independently of the
remaining blade elements.

2.08.2.3 The Blade-Element Momentum Theory


The BEM method was developed by Glauert [3] as a practical way to analyze and design rotor blades. In the BEM theory, the
loading is computed using two independent methods, that is, by a local blade-element consideration using tabulated
two-dimensional (2D) airfoil data and by use of the 1D momentum theorem. First, employing BEM, axial load and torque are
written as, respectively,

Table 1 Flow conditions for the optimum actuator disk

a a′ a′x2λ2 λx

0.25 ∞ 0 0
0.26 5.500 0.0296 0.073
0.27 2.375 0.0584 0.157
0.28 1.333 0.0864 0.255
0.29 0.812 0.1136 0.374
0.30 0.500 0.1400 0.529
0.31 0.292 0.1656 0.753
0.32 0.143 0.1904 1.150
0.33 0.031 0.2144 2.630
0.333 0.003 01 0.2216 8.58
1/3 0 0.2222 ∞

Table 2 Power coefficient as function


of tip speed ratio for optimum actuator
disk

λ CP max

0.5 0.288
1.0 0.416
1.5 0.480
2.0 0.512
2.5 0.532
5.0 0.570
7.5 0.582
10.0 0.593
Aerodynamic Analysis of Wind Turbines 229

z
Wi U∞

L
Vrel

φ α

γ D

–θ Vθ –Ωr

Figure 2 Cross-sectional airfoil element.

dT 1
¼ BFn ¼ ρcBVrel
2
⋅ Cn ½22
dr 2
dM 1
¼ BrFt ¼ ρcBrVrel
2
⋅ Ct ½23
dr 2
where c is the blade chord, B is the number of blades, Vrel is the relative velocity, Fn and Ft denote the loading on each blade in axial
and tangential direction, respectively, and Cn and Ct denote the corresponding 2D tabulated force coefficients.
From the velocity triangle at the blade element (see Figure 2), we deduce that

U∞ ð1− aÞ Ωr ð1 þ a′Þ
sin  ¼ ; cos  ¼ ½24
Vrel Vrel
where the induced velocity is defined as Wi = (−aU0, a′Ωr). Using the above relations, we get

U∞2 ð1 − a Þ 2 U∞ ð1− aÞΩrð1 þ a′Þ


2
Vrel ¼ ¼ ½25
sin 2  sin  cos 
Inserting these expressions into eqns [22] and [23], we get

dT ρBcU02 ð1− aÞ 2
¼ Cn ½26
dr 2 sin 2 

dM ρBcU0 ð1− aÞΩr 2 ð1 þ a′Þ


¼ Ct ½27
dr 2 sin  cos 
Next, applying axial momentum theory, the axial load is computed as

dT
¼ ρðU0 − uwake Þ2πruR ¼ 4πρU02 að1− aÞ ½28
dr
where uR = U0(1 − a) is the axial velocity in the rotor plane and uwake = U0(1 − 2a) is the axial velocity in the ultimate wake.
Applying the moment of momentum theorem, we get

dM
¼ ρruθ 2πruR ¼ 4πρr 3 ΩU0 a′ð1− aÞ ½29
dr
where uθ = 2Ωra′ is the induced tangential velocity in the far wake. Combining eqns [26] and [27] with eqns [28] and [29], we get
after some algebra
1
a¼ ½30
4 sin 2 =ðσCn Þ þ 1

1
a′¼ ½31
4 sin  cos =ðσCt Þ− 1

2.08.2.3.1 Tip correction


Since the above equations are derived assuming azimuthally independent stream tubes, they are only valid for rotors with infinite
many blades. In order to correct for finite number of blades, Glauert [3] introduced Prandtl’s tip loss factor. In this method, a
correction factor, F, is introduced that corrects the loading. In a recent paper by Shen et al. [9], the tip correction is discussed and
various alternative formulations are compared. However, here we limit the correction to the original form given by Glauert [3]. In
this model, the induced velocities are corrected by the tip loss factor F, modifying eqns [28] and [29] as follows,
230 Aerodynamic Analysis of Wind Turbines

dT
¼ 4πrρU∞2 aFð1− aÞ ½32
dr

dM
¼ 4πρr 3 ΩU∞ a′Fð1− aÞ ½33
dr
where σ = Bc/2πr. An approximate formula of the Prandtl tip loss function was introduced as follows,
  
2 BðR − rÞ
F¼ cos − 1 exp − ½34
π 2r sin 
where  = (r) is the angle between the local relative velocity and the rotor plane.
The coefficients (Cn, Ct) are related to the lift and drag coefficients (Cl, Cd) by Cn = Cl cos  + Cd sin  and Ct = Cl sin  − Cd cos ,
respectively. (Cl, Cd) depend on local airfoil shape and are obtained using tabulated 2D airfoil data corrected with 3D rotating
effects. Equating eqn [26] to eqn [32] and eqn [27] to eqn [33], the final expressions for the interference factors read
1
a¼ ½35
4F sin2 =ðσCn Þ þ 1

1
a′¼ ½36
4F sin  cos =ðσCt Þ−1

2.08.2.3.2 Correction for heavily loaded rotors


By putting eqn [32] into dimensionless form, we get the following expression for the local thrust coefficient,

dT
CT ¼ ¼ 4aFð1− aÞ ½37
2 ρU∞ 2πrdr
1 2

For heavily loaded rotors, that is, for a values between 0.3 and 0.5, this expression ceases to be valid as the wake velocity tends to
zero with an unrealistic large expansion as a result. It is therefore common to replace it by a simple empirical relation. Following
Glauert [3], an appropriate correction is to replace the expression for a ≥ 1/3 with the following expression:
 a
CT ¼ 4aF 1− ð5 − 3aÞ ½38
4
As discussed in, for example, Spera [10] or Hansen [11], other expressions can also be used.

2.08.2.3.3 Yaw correction


Yaw refers to the situation where the incoming flow is not aligned with the rotor axis. In this case, the wake flow is not in line with
the free wind direction and it is impossible to apply the usual control volume analysis. A way of solving the problem is to maintain
the control volume and specify an azimuth-dependent induction. In practice, it works by computing a mean induction and
prescribe a function that gives the azimuthal dependency of the induction. The following simple formula has been proposed by
Snel and Schepers [12],
 χ 
r
wi ¼ wi0 1 þ tan cos ðθblade − θ0 Þ ½39
R 2

where wi0 is the annulus averaged induced velocity and χ is the wake skew angle, which is not identical to the yaw angle because the
induced velocity in yaw alters the mean flow direction in the wake flow. In the notation used here, θblade denotes the azimuthal
position of the blade and θ0 is the azimuthal position where the blade is deepest in the wake. For more details, the reader is referred
to the text book by Hansen [11].

2.08.2.3.4 Dynamic wake


Dynamic wake or dynamic inflow refers to unsteady flow phenomena that affect the loading on the rotor. In a real flow
situation, the rotor is subject to unsteadiness from coherent wind gusts, yaw misalignment, and control actions, such as
pitching and yawing. When the flow changes in time, the wake is subject to a time delay when going from one equilibrium
state to another. An initial change creates a change in the distribution of trailing vorticity which then is convected
downstream and first can be felt in the induced velocities after some time. However, the BEM method in its simple form is
basically steady; hence, unsteady effects have to be included as an additional ‘add-on’. In the European CEC Joule II project
‘Dynamic Inflow: Yawed Conditions and Partial Span Pitch’ (see Schepers and Snel [13]), various dynamic inflow models were
developed and tested. Essentially, a dynamic inflow model predicts the time delay through an exponential decay with a time
constant corresponding to the convective time of the flow in the wake. As an example, the following simple model was
suggested,
Aerodynamic Analysis of Wind Turbines 231

 r du ΔT
i
Rf þ 4ui ðU0 − ui Þ ¼ ½40
R dt 2πrΔr
where the function f(r/R) is a semiempirical function associated with the induction. The equation can be seen to correspond to the
axial momentum equation, eqn [28], except for the time-term that is responsible for the time delay.

2.08.2.3.5 Airfoil data


As a prestep to the BEM computations, 2D airfoil data have to be established from wind tunnel measurements. In order to construct
a set of airfoil data to be used for a rotating blade, the airfoil data further need to be corrected for 3D and rotational effects. A simple
correction formula for rotational effects was proposed by Snel and van Holten [14] for incidences up to stall. For higher incidences
(>40°), 2D lift and drag coefficients of a flat plate can be used. These data, however, are too big because of aspect ratio effects and
here the correction formulas of Viterna and Corrigan [15] are usually applied (see also Spera [10]). Furthermore, since the angle of
attack is constantly changing due to fluctuations in the wind and control actions, it is needed to include a dynamic stall model to
compensate for the time delay associated with the dynamics of the boundary layer and wake of the airfoil. This effect can be
simulated by a simple first-order dynamic model, as proposed by Øye [16], or it can be considerably more advanced, taking into
account also attached flow, leading edge separation and compressibility effects, as in the model of Leishman and Beddoes [17].

2.08.3 Advanced Aerodynamic Modeling

Although the BEM method is widely used and today constitutes the only design methodology in use by industry, there is a big need
for more sophisticated models for understanding the underlying physics. Various numerically based aerodynamic rotor models
have in the past years been developed, ranging from simple lifting line wake models to full-blown Navier–Stokes-based CFD
models. In the following, the most used models will be introduced.

2.08.3.1 Vortex Models


Vortex wake models denote a class of methods in which the rotor blades and the trailing and shed vortices in the wake are
represented by lifting lines or surfaces. At the blades, the vortex strength is determined from the bound circulation which is related to
the local inflow field. The global flow field is determined from the induction law of Biot–Savart, where the vortex filaments in the
wake are advected by superposition of the undisturbed flow and the induced velocity field. The trailing wake is generated by
spanwise variations of the bound vorticity along the blade. The shed wake is generated by the temporal variations as the blade
rotate. Assuming that flow in the region outside the trailing and shed vortices is curl-free, the overall flow field can be represented by
the Biot–Savart law. Utilizing the Biot–Savart law, simple vortex models can be derived to compute quite general flow fields about
wind turbine rotors. The first example of a simple vortex model is the one due to Joukowsky [18], who proposed to model the wake flow
by a hub vortex plus tip vortices represented by an array of semi-infinite helical vortices with constant pitch (see also Margoulis [19]).
However, this model contains inherent problems due to the singular behavior of the vortices, and as an axisymmetric approximation,
one may represent the tip vortices as a series of ring vortices.
To compute flows about actual wind turbines, it becomes necessary to combine the vortex line model with tabulated 2D airfoil
data. This can be accomplished by representing the spanwise loading on each blade by a series of straight vortex elements located
along the quarter chord line. The strength of the vortex elements are determined by employing the Kutta–Joukowsky theorem on the
basis of the local airfoil characteristics. When the loading varies along the span of each blade, the value of the bound circulation will
change from one filament to the next. This is compensated for by introducing trailing vortex filaments whose strengths correspond
to the differences in bound circulation between adjacent blade elements. Likewise, shed vortex filaments are generated and advected
into the wake whenever the loading undergoes a temporal variation. While vortex models generally provide physically realistic
simulations of the flow structures in the wake, the quality of the obtained results still depends on the input airfoil data.
In vortex models, the flow structure can either be prescribed or computed as a part of the overall solution procedure. In a
prescribed vortex technique, the position of the vortical elements is specified from measurements or semiempirical rules. This makes
the technique fast to use on a computer, but limits its range of application to more or less well-known steady flow situations. For
unsteady flow situations and complicated flow structures, free wake analysis becomes necessary. A free wake method is more
straightforward to understand and use, as the vortex elements are allowed to advect and deform freely under the action of the
velocity field. The advantage of the method lies in its ability to calculate general flow cases, such as yawed wake structures and
dynamic inflow. The disadvantage, on the other hand, is that the method is far more computing expensive than the prescribed wake
method, since the Biot–Savart law has to be evaluated for each time step taken. Furthermore, free-vortex wake methods tend to
suffer from stability problems owing to the intrinsic singularity in induced velocities that appears when vortex elements are
approaching each other. This can to a certain extent be remedied by introducing a vortex core model in which a cut-off parameter
models the inner viscous part of the vortex filament. In recent years, much effort in the development of models for helicopter rotor
flow fields have been directed toward free wake modeling using advanced pseudo-implicit relaxation schemes, in order to improve
numerical efficiency and accuracy (see Leishman [20]). A special version of the free-vortex wake methods is the method by
Voutsinas [21] in which the flow modeling is taken care of by vortex particles or vortex blobs.
232 Aerodynamic Analysis of Wind Turbines

A generalization of the vortex method is the so-called Boundary Integral Element Method (BIEM). Where the rotor blade in a
simple vortex method is represented by straight vortex filaments, the BIEM takes into account the actual finite thickness geometry of
the blade. The theoretical background for BIEMs is potential theory where the flow, except at solid surfaces and wakes, is assumed to
be irrotational. In a rotor computation, the blade surface is covered with both sources and doublets, while the wake only is
represented by doublets (see, e.g., Katz and Plotkin [22] or Cottet and Koumoutsakos [23]). The circulation of the rotor is obtained
as an intrinsic part of the solution by applying the Kutta condition on the trailing edge of the blade. The main advantage of the BIEM
is that complex geometries can be treated without any modification of the model. Thus, both the hub and the tower can be modeled
as a part of the solution. Furthermore, the method does not depend on airfoil data and viscous effects can, at least in principle, be
included by coupling the method to a viscous solver.

2.08.3.2 Numerical Actuator Disk Models


The actuator disk denotes a technique for analyzing rotor performance. In this model, the rotor is represented by a permeable disk
that allows the flow to pass through the rotor, at the same time as it is subject to the influence of the surface forces. The ‘classical’
actuator disk model is based on conservation of mass, momentum, and energy, and constitutes the main ingredient in the 1D
momentum theory. Combining it with a blade-element analysis, we end up with the BEM model. In its general form, however, the
actuator disk might as well be combined with a numerical solution of the Euler or Navier–Stokes equations.
In a numerical actuator disk model, the Navier–Stokes (or Euler) equations are typically solved by a second-order accurate finite
difference/volume scheme, as in a usual CFD computation. However, the geometry of the blades and the viscous flow around the
blades are not resolved. Instead, the swept surface of the rotor is replaced by surface forces that act upon the incoming flow. This can
either be implemented at a rate corresponding to the period-averaged mechanical work that the rotor extracts from the flow or by
using local instantaneous values of tabulated airfoil data. In the simple case of an actuator disk with constant prescribed loading,
various fundamental studies can easily be carried out. The generalized actuator disk method resembles the BEM method in the sense
that the aerodynamic forces has to be determined from measured airfoil characteristics, corrected for 3D effects, using a
blade-element approach. For airfoils subjected to temporal variations of the angle of attack, the dynamic response of the
aerodynamic forces changes the static aerofoil data and dynamic stall models have to be included. The first computations of
wind turbines employing numerical actuator disk models in combination with a blade-element approach were carried out by
Sørensen and Myken [24] and Sørensen and Kock [25]. This was later followed by different research groups who employed the
technique to study various flow cases, including coned and yawed rotors, rotors operating in enclosures, and wind farm simulations.
For a review on the method, the reader is referred to Vermeer et al. [26], Hansen et al. [27], or the VKI Lecture Series [28].
The main limitation of the axisymmetric assumption is that the forces are distributed evenly along the actuator disk; hence, the
influence of the blades is taken as an integrated quantity in the azimuthal direction. To overcome this limitation, an extended 3D
actuator disk model has been developed by Sørensen and Shen [29]. The model combines a 3D Navier–Stokes solver with a
technique in which body forces are distributed radially along each of the rotor blades. Thus, the kinematics of the wake flow is
determined by a full 3D Navier–Stokes simulation, whereas the influence of the rotating blades on the flow field is included using
tabulated airfoil data to represent the loading on each blade. As in the axisymmetric model, airfoil data and subsequent loading are
determined iteratively by computing local angles of attack from the movement of the blades and the local flow field. The concept
enables one to study in detail the dynamics of the wake and the tip vortices and their influence on the induced velocities in the rotor
plane. A model following the same idea has been suggested by Leclerc and Masson [30]. A main motivation for developing such
types of model is to be able to analyze and verify the validity of the basic assumptions that are employed in the simpler more
practical engineering models. Reviews of the basic modeling of actuator disk and actuator line models can be found in the PhD
dissertations of Mikkelsen [31], Troldborg [32], and Ivanell [33].

2.08.3.3 Full Navier–Stokes Modeling


During the past four decades, a strong research activity within the aeronautical field has resulted in the development of a series of
CFD tools based on the solution of the Navier–Stokes equations. Within aerodynamics, this research has mostly been related to
flows around fixed-wing aircraft and helicopters. Looking specifically on the aerodynamics of horizontal-axis wind turbines, we find
some striking differences as compared to usual aeronautical applications. First, as tip speeds generally never exceed 100 m s−1, the
flow around wind turbines is incompressible. Next, the optimal operating condition for a wind turbine always includes stall, with
the upper side of the rotor blades being dominated by large areas of flow separation. This is in contrast to the cruise condition of an
aircraft where the flow is largely attached.
Some of the experience gained from the aeronautical research institutions has been exploited directly in the development of CFD
algorithms for wind turbines. Notably is the development of basic solution algorithms and numerical schemes for solution of the
flow equations, grid generation techniques, and the modeling of boundary layer turbulence. These elements together form the basis
of all CFD codes, of which some already have existed for a long time as standard commercial software.
Today, there exist two main paths to follow when conducting CFD computations; either the equations are solved by using
Reynolds averaging or by introducing space filtration. The most popular method is based on solving the Reynolds-averaged
Navier–Stokes (RANS) equations, closing the system by introducing a suitable one-equation or two-equation turbulence model,
such as the Spalart–Allmaras [34] or the k − ε [35] model. By using this kind of model, only the time-averaged flow field is
Aerodynamic Analysis of Wind Turbines 233

computed, whereas the unsteady field is modeled through the turbulence model. If the flow is dominated by a broad spectrum of
time scales, the low frequencies may be simulated partly by maintaining the time-term in the RANS equations. In this case, it is
sometimes referred to as URANS (unsteady RANS). The advantage of RANS or URANS is that a fully resolved computation can be
established with some few million mesh points, which makes it possible to reach a full 3D solution even on a portable computer.
In the past years, refined one- and two-equation turbulence models have been developed to cope with specific flow features.
In particular, the k − ω SST model developed by Menter [36] has shown its capability to cope with lightly separated airfoil flows, and
today this model is widely used for wind turbine computations. The accuracy of the computations, however, is restricted by the
turbulence model’s lack of ability of representing a full unsteady spectrum. Thus, for attached flow the accuracy is fully adequate,
whereas for stalled flows, it may degenerate completely. This is further rendered complicated by the laminar–turbulent transition
process that also has to be modeled in order compute the onset of turbulence. An alternative to RANS/URANS is LES. In LES, the
Navier–Stokes equations are filtered spatially on the computational mesh and only the subgrid scale (SGS) part of the turbulence is
modeled using a so-called SGS model. The advantage of LES is that all the dynamics of the flow field is captured and that accurate
solutions can be obtained even under highly separated flow conditions. The computational price, however, is often prohibitive,
even when solving parallelized computing algorithms on large cluster systems, because of the large number of mesh points that are
needed to resolve practical flows at high Reynolds numbers. As compared to direct numerical simulation (DNS), where the
Navier–Stokes equations are solved directly without any modeling of the turbulence, LES is, however, still several orders of
magnitude faster.
To give an estimate of computing expenses and the number of mesh points required to resolve a turbulent flow field, one can use
the Kolmogorov length scale, ℓ, as the smallest scale and the length of the considered object, L, as the largest length scale. According to
3=4
Lesieur et al. [37], an estimate of the ratio between the largest and the smallest length scale can be given as L=ℓ ≈ ReL , where the
Reynolds number ReL = UL/υ, with U denoting a characteristic wind speed and υ is the kinematic viscosity. For an airfoil of a wind
turbine blade, a typical value is ReC ≅ 5⋅106 where index C denotes the chord length. Thus, for a DNS computation of an airfoil section,
we need in the order of 105 mesh points in each direction, resulting in a total of approximately 1015 mesh points. For a corresponding
LES computation, this may be reduced to about 1010 mesh points, if we assume that the SGS covers about 1.5 decades. A main
difference between RANS and LES is that RANS computations may be carried out in a pure 2D domain, for example, when studying or
designing airfoils, whereas LES is always intrinsically unsteady and 3D. As a compromise between the fast computing time of RANS
methods and the accuracy of LES, Spalart et al. [38] developed the detached eddy simulation (DES) technique. This technique is a
hybrid approach in which the flow near boundaries is solved using a traditional RANS turbulence model and the outer flow is
modeled using a SGS model. However, this technique puts severe bounds on the grid, since very high aspect ratios are needed near the
boundaries, whereas the grid is required to be as isotropic as possible in the LES domain. When computing wakes, the number of mesh
points need not depend on the Reynolds number, if for example, the influence from the surface is ignored. In this case most of the flow
can be simulated by using LES technique to simulate the dynamics of the main vortex structures and model the smaller scales by an
SGS model. However, if one wishes to include the surface-bounded boundary layer in the computation the number of mesh points is
mainly determined by the Reynolds number, which for a modern wind turbine of a diameter of about 100 m is about ReD ≅ 107. An
overview of the required number of mesh points for different approaches is given in Table 3.

2.08.4 CFD Computations of Wind Turbine Rotors

The research on CFD in wind turbine aerodynamics was initiated through European Union-sponsored collaborate projects,
such as VISCWIND [39], VISCEL [40], and KNOW-BLADE in Europe. The first full Navier–Stokes simulation for a complete
rotor blade was carried out by Sørensen and Hansen [41] and later followed by Duque et al. [42] and Sørensen et al. [43] in
connection with the American NREL experiment at NASA Ames and the accompanying National Renewable Energy Laboratory/
National Wind Technology Center (NREL/NWTC) aerodynamics blind comparison test [44]. This experiment has achieved a
significant new insight into wind turbine aerodynamics and revealed serious shortcomings in present-day wind turbine
aerodynamics prediction tools. First, computations of the performance characteristics of the rotor by methods based on the
BEM technique were found to be extremely sensitive to the input blade section aerodynamic data. The predicted values of the
distribution of the normal force coefficient deviated from measurements by as much as 50%. Even at low angles of attack,
model predictions differed from measured data by 15–20% [44]. Next, the computations based on Navier–Stokes equations
convincingly showed that CFD had matured to become an important tool for predicting and understanding the flow physics of

Table 3 Number of required mesh points for


various types of computations

Airfoil Full rotor Wake

RANS 105 107 105 – 108

DES 107 108 107 – 1010

LES 1010 1012 107 – 1014

DNS 1015 1017 107 – 1019

234 Aerodynamic Analysis of Wind Turbines

Wind turbine blade, 19.1 meter, suction side


Blade root
Leading edge
Separation line

Wind
Reattchment line

Trailing edge

tion
Blade revolu Anomalous vortex Blade tip

Windspeed 15 m/s, 27 RPM

Figure 3 Sketch of flow topology and limiting streamlines on a wind turbine blade.

modern wind turbine rotors. The Navier–Stokes computations by Sørensen et al. [43] generally exhibited good agreement with
the measurements up to wind speeds of about 10 m s−1. At this wind speed, flow separation sets in and for higher wind speeds
it dominates the boundary layer characteristics. Hence, it is likely that the introduction of a more physically consistent
turbulence modeling and the inclusion of a laminar/turbulent transition model will improve the quality of the results
(Sørensen [45]). A large number of full 3D Navier–Stokes computations have later been carried out by different research
groups. The computations include RANS and DES simulations of full rotor systems, the hub, studies of tip flows, blade–tower
interaction, and wind turbine blades under parked conditions. Reviews can be found in Hansen et al. [27] and Sørensen [46],
and various contributions were published in the proceedings from TWIND2007 [47]. To illustrate the degree of complexity one
obtains using a full 3D Navier–Stokes methodology in Figure 3, we show a computation of a rotating 19.1 m long wind
turbine blade. It is clearly seen here that a complicated flow topology results, including a large separated area, which could not
be obtained using the BEM technique or inviscid computations.

2.08.5 CFD in Wake Computations

Modern wind turbines are often clustered in wind parks in order to reduce the overall installation and maintenance expenses. Because
of the mutual interference between the wakes of the turbines, the total power production of a park of wind turbines is reduced as
compared to an equal number of stand-alone turbines. Thus, the total economic benefit of a wind park is a trade-off between the
various expenses to erect and operate the park, the available wind resources at the site, and the reduced power production because of
the mutual influence of the turbines. A further unwanted effect is that the turbulence intensity in the wake is increased because of the
interaction from the wakes of the surrounding wind turbines. As a consequence, dynamic loadings are increased that may excite the
structural parts of the individual wind turbine and enhance fatigue loadings. The turbulence created from wind turbine wakes is mainly
due to the dynamics of the vortices originating from the rotor blades. The vortices are formed as a result of the rotor loading. To analyze
the genesis of the wake, it is thus necessary to include descriptions of the aerodynamics of both the rotor and the wake. Although many
wake studies have been performed over the past two decades, a lot of basic questions still need to be clarified in order to elucidate the
dynamic behavior of individual as well as multiple interactive wakes behind wind turbines.
When regarding wakes, a distinct division can be made between the near- and the far-wake region. The near wake is normally
taken as the area just behind the rotor, where the properties of the rotor can be discriminated, so approximately up to 1 rotor
diameter downstream. Here, the presence of the rotor is apparent by the number of blades, blade aerodynamics, including stalled
flow, 3D effects, and the tip vortices. The far wake is the region beyond the near wake, where the focus is put on the influence of
wind turbines in park situations; hence, modeling the actual rotor is less important. The near wake research is focused on the
performance and the physical process of power extraction, while the far wake research is more focused on the mutual influence
when wind turbines are placed in clusters or wind farms.
The far wake has been a subject of extensive research both experimentally and numerically. Semianalytical far wake models have
been proposed to describe the wake velocity after the initial expansion (e.g., Ainslie [48]). Detailed numerical studies of the far wake
have been carried out by Crespo and Hernández [49] using methods based on the UPMWAKE model in which the wind turbine is
supposed to be immersed in an atmospheric boundary layer. This model uses a finite difference approach and a parabolic
approximation to solve the RANS equations combined with a k − ε turbulence model.
As illustrated in Table 3, prohibitively many mesh points are needed if one wishes to carry out LES or DNS of the wake in an
atmospheric boundary layer. However, employing the actuator line technique and representing the ambient turbulence and shear
Aerodynamic Analysis of Wind Turbines 235

flow by body forces, the number of mesh points become affordable even for a high Reynolds number LES computation. Using
this technique, near-wake computations have been carried out by Sørensen and Shen [29], Ivanell et al. [50, 51], and Troldborg et al.
[52, 53]. In a recent survey by Vermeer et al. [26], both near-wake and far-wake aerodynamics are treated, whereas a survey focusing
solely on far-wake modeling was earlier given by Crespo et al. [54].
To illustrate the type of results that may be achieved combining LES and the actuator line technique, results obtained from
simulations by Troldborg et al. [53] of a stand-alone wind turbine will be presented. The computations were carried out using airfoil
data from the Tjæreborg wind turbine. The blade radius of this turbine is 30.56 m and it rotates at 22.1 rpm, corresponding to a tip
speed of 70.7 m s−1. The blade sections consist of NACA 44xx airfoils with a chord length of 0.9 m at the tip, increasing linearly to
3.3 m at hub radius 6 m. The blades are linearly twisted 1° per 3 m.
Figure 4 shows instantaneous vortex structures in the near wake of a rotor operating at a wind speed of W0 = 6 m s−1,
corresponding to a tip speed ratio of about 12. It is seen here that the wake flow collapses into small-scale turbulence about 1
diameter behind the rotor. The actual position depends on the loading and the ambient turbulence level. In the simulation in
Figure 4 the inflow conditions were pure laminar and the collapse is due to intrinsic instabilities of the flow. Figure 5 shows the
contours of the instantaneous absolute vorticity in the x/R = 9 plane for three different cases. Regions of high vorticity appear as light
colors. Note that the rotor is located to the left in the plots and that only the downstream development of the wake is shown. The

Figure 4 Vortex structures in the near wake after the Tjæreborg wind turbine (Troldborg et al. [53]).

Figure 5 Downstream development of wake behind wind turbine. Upper figure: W0 = 6 m s−1; middle figure: W0 = 10 m s−1; lower figure: W0 = 14 m s−1.
In all figures, the rotor is located to the left (Troldborg et al. [53]).
236 Aerodynamic Analysis of Wind Turbines

bound vorticity of the blades is seen to be shed downstream from the rotor in individual vortex tubes. A closer inspection of the
vorticity contours at W0 = 6 m s−1 revealed that the distinct tip-vortex pattern is preserved about 0.5 rotor radii downstream, where
after they smear into a continuous vorticity sheet. In the case where the free stream velocity is W0 = 10 and 14 m s−1, distinct tip
vortices can be observed about 1.5 and 7 rotor radii downstream, respectively. In all cases, the structures might have been preserved
even further if a finer grid had been used. Moreover, it should be noted that using the absolute value of the vorticity as a means of
identifying vortices is limited by its strong dependence on the chosen contour levels, and therefore, vortex structures might very well
be present even though they are not immediately visible. For the rotor operating at the highest tip speed ratio, instability of the tip
vortices are observed only 2 and 5 rotor radii downstream where the entire wake flow completely breaks up. In the case where
W0 = 10 m s−1, the tip vortices are observed to undergo a Kelvin–Helmholtz instability approximately 7 rotor radii downstream. The
root vortex also becomes unstable at this position. Further downstream the root and tip vortices interact, which causes the flow to
become fully turbulent. Instability of the tip vortices is also observed in the last case where W0 = 14 m s−1, but as expected it takes
place even further downstream (ca. 10 rotor radii downstream) and is not as strong due to the generally higher stability and
persistence of the tip vortices, when the tip speed ratio and thus also the thrust is low.

2.08.6 Rotor Optimization Using BEM Technique

In the past three decades, the size of commercial wind turbines has increased from units of about 50 kW in the early 1980s to the latest
multi-MW turbines with rotor diameters over 120 m. In spite of repeated predictions of a leveling off at an optimum mid-range size
and periods of stagnation, the size of commercial wind turbines has steadily increased with about a 5-doubling in installed generator
power over a period of one decade. The overall goal is to reduce the cost price of the produced energy, and as long as increasing the size
results in a reduction of the cost price, it is likely that the wind turbines will increase in size also for many years to come. There are
obviously factors that may bring this trend to an end, such as problems related to the handling and manufacturing of the large blades.
A more sophisticated way of capturing more energy from the wind, however, is to improve the aerodynamic efficiency of the energy
conversion by using optimization techniques in the initial design. In the development of new wind turbines, aerodynamic and
structural optimization has become an important issue for optimizing the energy yield and thereby minimizing the cost price of the
produced energy. How to reduce the cost of a wind turbine per unit of energy is an important task in modern wind turbine research.
Classical models for aerodynamic optimization of rotors can be found in the text books of Glauert [3] and Theodorsen [55] and in
revised form by Okulov and Sørensen [56, 57]. However, an aerodynamic optimal rotor may not necessarily be the most cost-effective,
as the target is to reduce the price of the produced energy. Since an optimization technique works together with aerodynamic and
structural models, results from an optimization procedure will often be influenced a great deal by the models used. Thus, accurate and
efficient models for predicting wind turbine performance are essential for obtaining reliable optimum designs of wind turbine rotors.
The first multidisciplinary optimization method for designing horizontal-axis wind turbines is due to Fuglsang and Madsen [58]. The
objective used in their method was to minimize the cost of energy employing multiple constrains. Generally, multiobjective
optimization methods are employed in which the blades are optimized by varying blade structural parameters such as stiffness,
stability, and material weight. Site specifics from sites comprising normal flat terrain, offshore, and complex terrain wind farms can
also be incorporated in the design process of the wind turbine rotors [59].
To illustrate the basics of design optimization of wind turbine rotors, in the following we show the features of an optimization
model developed at Technical University of Denmark and ChongQing University [60] for optimizing the geometry of wind turbines
to maximize the energy yield. The method is based on combining an aeroelastic model containing 11 degrees of freedom with the
BEM technique. The most important issue when performing optimizations is to locate the main parameters and a suitable object
function. In the model, the object function is defined as the minimum cost price of the produced energy, determined by computing
the annual energy production (AEP) and the production cost of the turbine. In the following, the cost model and the design
variables used in the optimization model are presented. As design variables we choose chord length, twist angle, relative thickness,
and tip pitch angle. Estimating the cost of a wind turbine is an important and difficult task, but also crucial for the success of an
optimization. The cost model includes the capital costs from foundation, tower, rotor blades, gearbox, and generator plus the costs
from operation and maintenance. The total cost of a wind turbine can be expressed as

X
N X
N
C¼ Ci ¼ Ri ðbi þ ð1− bi Þwi Þ ½41
i ¼1 i ¼1

where Ci is the cost of the i-th component of the wind turbine and N is the number of main components, Ri is the initial cost of the
i-th component determined from a reference rotor, bi is the fixed part of the i-th component that counts for manufacturing and
transport, (1 − bi) is the variable part of the i-th component, and wi is the weight parameter of the i-th component. The weight
parameter in Fuglsang and Madsen [58] was dependent on the design loads of extreme forces and moments and lifetime equivalent
fatigue forces and moments. To get more information about the cost of a whole wind turbine, the reader is referred to Fuglsang and
Madsen [58]. As the costs from operation and maintenance often can be counted as a small percentage of the capital cost, reduction
of the capital cost becomes the essential task for the design. Further, a well-designed wind turbine with a low energy cost always has
an aerodynamically efficient rotor. Therefore, the rotor design plays an important role for the whole design procedure of a wind
turbine. In the current study, we restrict our objective to the cost from the rotor. Thus, the objective function is defined as,
Aerodynamic Analysis of Wind Turbines 237

Crotor
f ðxÞ ¼ COE ¼ ½42
AEP
where COE is the cost of energy of a wind turbine rotor and Crotor is the total cost for producing, transporting, and erecting a wind
turbine rotor. In the current study, the fixed part of the cost for a wind turbine rotor brotor is chosen to be 0.1. Therefore, the total cost
of a rotor, Crotor, is a relative value defined as

Crotor ¼ brotor þ ð1− brotor Þwrotor ½43


where wrotor is the weight parameter of the rotor. In the present study, the weight parameter is calculated from the chord and mass
distributions of the blades. Dividing a blade into n cross-sections, wrotor is estimated as
X n
mi ⋅ ci ; opt
wrotor ¼ ½44
i ¼1
Mtot ⋅ ci ; orig

where mi is the mass of the i-th cross-section of the blade, ci,opt is the mean chord length of the i-th cross-section of the optimized blade,
ci,orig is the mean chord length of the i-th cross-section of the original blade, and Mtot is the total mass of the blade. The power curve is
determined from the BEM method. In order to compute the AEP, it is necessary to combine the power curve with the probability density
of wind speed (i.e., the Weibull distribution). The function defining the probability density can be written in the following form
 k !  !
Vi Vi þ 1 k
f ðVi < V < Vi þ 1 Þ ¼ exp − − exp − ½45
A A

where A is the scale parameter, k is the shape factor, and V is the wind speed. In the current study, the shape factor is chosen to be
k = 2, corresponding to the Rayleigh distribution. If a wind turbine operates the full 8760 h yr−1, its AEP is computed as

X
M −1
1
AEP ¼ ðPðViþ1 Þ þ P ðV Þ i Þ  f ðVi < V < Viþ1 Þ  8760 ½46
i¼1
2

where P(Vi) is the power at wind speed Vi and M denotes the number of wind speeds considered.
As an example of the optimization model, we here show how the performance of the Tjæreborg 2 MW rotor may be improved
using optimization. The optimization is based on the original rotor; thus, the rotor diameter and the rotational speed are chosen to
be the same, whereas chord length, twist angle, relative thickness, and tip pitch angle are chosen as design variables. A cubic
polynomial is used to control the chord distribution and a spline function is used to control the distributions of twist angle and
relative thickness. Since the cost of a rotor depends on the lifetime of the blades, power output, shaft torque, and thrust are
constrained in the optimization process. The values are here constrained not to exceed the values of the original design. As a usual
procedure for optimization problems, we have one objective function and multiple constraints. To achieve the optimization, the
fmincon function in Matlab is used.
The Tjæreborg turbine is equipped with a three-bladed rotor of radius 30.56 m. In the BEM computations, 20 uniformly
distributed blade elements are used. The optimization design is performed from a radial position at a radius of 6.46 m to the tip
of the blade. In the optimization process, the lower limits for chord, twist angle, and relative thickness are 0 m, 0°, and 12.2%,
respectively, and the upper limits are 3.3 m, 8°, and 100%, respectively. To reduce the computational time, four points along the
blade are used to control the shape of the blade. The outcome of the optimization is shown in Figure 6, in which the chord and twist
distributions of the original and the optimized Tjæreborg rotor are compared.

(a) 3.5 (b)10


Optimized rotor Optimized rotor
3 Orginal rotor Orginal rotor
8
2.5
Twist angle (°)
Chord (m)

6
2

1.5 4
1

0.5

0 0
5 10 15 20 25 30 5 10 15 20 25 30
Radius (m) Radius (m)

Figure 6 (a) Chord and (b) twist angle distributions of the original and the optimized Tjæreborg 2 MW rotor. Reproduced from Xudong W, Shen WZ, Zhu
WJ, et al. (2009) Shape optimization of wind turbine blades. Wind Energy 12(8): 781–803.
238 Aerodynamic Analysis of Wind Turbines

From Figure 6(a), it is seen that the optimized blade attains a remarkable reduction in chord length in the region between 10
and 23 m, as compared to the original rotor. At a radius of 15 m, the chord reduction reaches a maximum value of about 16%.
From a position at a radius of 23 m to a position at a radius of 28 m, the optimized chord has almost the same value as the
original distribution of the chord, whereas it decreases significantly in the tip region. The twist angle, Figure 6(b), is slightly
smaller than the original distribution. The performance of the optimized rotor is computed using the aerodynamic/aeroelastic
code and compared to the original rotor. Since the Tjæreborg rotor is a pitch-controlled rotor, the output power of the rotor is set
to be the rated power of 2 MW when the wind speed is larger than the rated wind speed of 15 m s−1. The AEP of the optimized
rotor is reduced about 4%, whereas the cost of the optimized rotor is reduced by about 7.1%. Thus, the cost of energy of the
Tjæreborg rotor is reduced about 3.4%.

2.08.7 Noise from Wind Turbines

Although offshore wind energy is evolving fast, most wind turbines are still placed in rural environments, where wind turbine noise
is of great concern since it may be the only major noise source. Machinery noise is generally not as important as aerodynamic noise,
as it has been reduced efficiently by well-known engineering techniques, such as proper insulation of the nacelle. As a rule of thumb,
aerodynamic noise from a wind turbine blade increases with the fifth power of the relative wind speed, as seen from the moving tip
of the blades. With the tip speed being the most significant parameter, aerodynamic noise has been controlled by lowering the tip
speed to a maximum of about 60 m s−1. However, in recent years, the biggest development of wind turbines has taken place
offshore, with the result that the latest generation of wind turbines operate at tip speeds up to 80 m s−1. Thus, for turbines erected
near the shore or for offshore turbines tested at land sites, noise has again become a subject of great concern with respect to public
acceptance. This is best illustrated by the increasing number of conferences concerning wind turbine noise; for example, in 2005 the
Initiative for Noise Control Engineering in Europe (INCE/Europe) initiated a biannual conference series on wind turbine noise,
which in 2009 took place in Aalborg, and in 2011 is scheduled to take place in Rome.
Through the years, several models have been proposed to explain and predict wind turbine noise. Some of the models are
somewhat simplistic, whereas others make use of complex CFD solvers that have not yet matured to be applied to compute noise
emission for realistic rotors (see Wagner et al. [61] for a thorough review of various models). As a compromise between computing
speed and accuracy, the most commonly used models are based on semiempirical relations. As a basis, most models employ the
experimental results and scaling laws on airfoil self-noise by Brooks et al. [62] together with the turbulence inflow model proposed
by Amiet [63]. This includes, for example, the models of Fuglsang and Madsen [64] and Zhu et al. [65, 66] and the model employed
in the SIROCCO project [67], as well as further developments by Moriarty et al. [68] and Lutz et al. [69].
In the following, we show some of the features of a typical semiempirical model, such as those referred to above. In the model,
only aerodynamic noise is considered (i.e., mechanical noise is not considered). Aerodynamic noise can be divided into ‘airfoil self-
noise’ and ‘turbulence inflow noise’. The former is a result of the interaction of the boundary layer of the airfoil with the trailing edge
and the latter results from the interaction of the existing turbulence in the wind with the airfoil. In the model, the airfoil self-noise
prediction is based on the functions given by Brooks et al. [62]. In total, five airfoil self-noise mechanisms were identified and
studied separately:

• turbulent boundary layer trailing edge noise,


• separation-stall noise,
• laminar boundary layer vortex shedding noise,
• tip vortex formation noise, and
• trailing edge bluntness vortex shedding noise.

As a result, scaling laws were proposed, yielding the sound pressure level at the observer position as a function of frequency for the
1/3 octave band spectrum. The scaling laws for the different mechanisms are all of similar form:
  
δi Mfði Þ LD
SPLi ¼ 10 log h
þ Fi ðStÞ þ Gi ðReÞ ½47
r2

where δi is the boundary layer displacement thickness, M is the Mach number, f(i) is the raised power that depends on
 is a sound directivity function, and r is the distance
the particular noise mechanism i, L is the airfoil section semi-span, Dh
to the observer. The additional terms Fi(St) and Gi(Re) are functions of the ‘Strouhal number’ St = fδ*/U and the ‘Reynolds
number’ Re. The nature of dependency is different for each noise mechanism but it is impressive that all the formulas look so
much alike.
For turbulent inflow, the prediction equation is normally based on the work of Amiet [63]. This model takes the following
form:
  − 7=3   
ΔL Kc
Lp ¼ 10 log ρ20 c02 l 2 M3 I2 ^k3 1 þ ^k 2 þ 58:4 þ 10 log ½48
r 1 þ Kc
Aerodynamic Analysis of Wind Turbines 239

where l is turbulence length scale, I is turbulence intensity, ρ0 is density, c0 is speed of sound, ΔL is blade segment semi-span, k^ is
corrected wave length, and Kc is low-frequency correction.
Taking into account all the variable dependencies, the problem of predicting the noise spectrum at a given observer position for a
given airfoil reduces to identifying the following quantities:

• The boundary layer thickness δ* at the trailing edge of the airfoil


• The relative wind speed defining M and Re
• The boundary layer transition type (forced or natural), leading to tripped or untripped flow
• Miscellaneous input parameters to the turbulence inflow noise model, such as turbulent length scale and intensity, in the model
reduced to the knowledge of the height from the ground z and the roughness length zo.

Here we do not go into the theory behind the empirical correlations, and for details about the nature of each of the modeled
noise mechanisms, we refer the reader to the original work of Brooks et al. [62] and Amiet [63] or the text book of Wagner
et al. [61].
As mentioned above, an important parameter for the calculation of airfoil self-noise is the boundary layer thickness at the
trailing edge. This can be calculated by use of the viscous–inviscid interactive computing program XFOIL [70]. It is important
to note that the scaling laws shown above are deduced from experiments based only on the NACA 0012 airfoil. For this
reason, an independent calculation of δ* for each airfoil type is vital. This was carried out for different values of the Reynolds
number and angle of attack and the computed boundary layer thickness was stored in a database and subsequently
determined by interpolation.
Essentially, the code consists of a ‘traditional’ BEM code, to compute the relative velocities along each blade element defining the
rotor, coupled with the routines to predict the noise contribution for each noise source along the span of the rotor blades. In short,
the prediction code works as follows. First, the relative velocities seen by the blade elements are computed, just like in an ordinary
BEM computation. Next, a table looking up in the boundary layer thickness database is made and the sound pressure level Lp and
the noise spectrum at the observer position are calculated for each noise mechanism and for each blade element. Finally, the sound
pressure levels are added for all elements, all blades, and all mechanisms and converted to sound power levels Lw referring to the
hub of the wind turbine.
The main advantage of the semiempirical model is that it is fast to run, even on a PC, and that it gives a surprisingly reliable
result. It is also fairly easy to couple the prediction code to an optimization algorithm and use it as a tool to optimize the rotor with
respect to both performance and noise.

References

[1] Lanchester FW (1915) A contribution to the theory of propulsion and the screw propeller. Transactions of the Institution of Naval Architects 57: 98.
[2] Betz AD (1920) Maximum der theoretisch möglichen Ausnützung des Windes durch Windmotoren. Zeitschrift für das gesamte Turbinenwesen 26: 307–309.
[3] Glauert H (1935) Airplane propellers. Division L. In: Durand WF (ed.) Aerodynamic Theory, vol. IV, pp. 169–360. Berlin, Germany: Springer.
[4] Rankine WJM (1865) On the mechanical principles of the action of propellers. Transactions of Institution of Naval Architects 6: 13.
[5] Froude RE (1889) On the part played in propulsion by difference of fluid pressure. Transactions of Institution of Naval Architects 30: 390–405.
[6] Froude W (1878) On the elementary relation between pitch, slip and propulsive efficiency. Transactions of Institution of Naval Architects 19: 47.
[7] van Kuik GAM (2007) The Lanchester-Betz-Joukowsky limit. Wind Energy 10: 289–291.
[8] Wilson RE and Lissaman PBS (1974) Applied Aerodynamics of Wind Power Machines. Corvallis, OR: Oregon State University.
[9] Shen WZ, Mikkelsen R, Sørensen JN, and Bak C (2005) Tip loss corrections for wind turbine computations. Wind Energy 8(4): 457–475.
[10] Spera DA (1994) Wind Turbine Technology. New York, NY: ASME Press.
[11] Hansen MOL (2008) Aerodynamics of Wind Turbines. London, UK: Earthscan.
[12] Snel H and Schepers JG (1994) Joint investigation of dynamic inflow effects and implementation of an engineering method. ECN-C-94-107, Netherlands Energy Research
Foundation ECN, Petten, The Netherlands.
[13] Schepers JG and Snel H (1995) Dynamic inflow: Yawed conditions and partial span pitch control. ECN-C-95-056, Netherlands Energy Research Foundation ECN, Petten, The
Netherlands.
[14] Snel H and van Holten T (1995) Review of recent aerodynamic research on wind turbines with relevance to rotorcraft. AGARD Report CP-552, ch. 7, pp. 1–11, AGARD Advisory
Group for Aerospace Research & Development.
[15] Viterna LA and Corrigan RD (1981) Fixed-Pitch Rotor Performance of Large HAWT’s. DOE/NASA Workshop on Large HAWTs. National Aeronautics and Space Administration,
Cleveland, Ohio.
[16] Øye S (1991) Dynamic stall, simulated as a time lag of separation. Proceedings of 4th IEA Symposium on the Aerodynamics of Wind Turbines. 20–21 November 1990, ETSU-N­
118. Harwell, UK.
[17] Leishman JG and Beddoes TS (1989) A semi-empirical model for dynamic stall. Journal of American Helicopter Society 34(3): 3–17.
[18] Joukowsky NE (1912) Vortex theory of a rowing screw. Trudy Otdeleniya Fizicheskikh Nauk Obshchestva Lubitelei Estestvoznaniya 16(1): 1–31.
[19] Margoulis W (1922) Propeller theory of Professor Joukowski and his pupils. NACA Technical Memorandum No. 79.
[20] Leishman JG (2002) Challenges in modelling the unsteady aerodynamics of wind turbines. Wind Energy 5: 85–132.
[21] Voutsinas SG (2006) Vortex methods in aeronautics: How to make things work. International Journal of Computational Fluid Dynamics 20(1): 3–18.
[22] Katz J and Plotkin A (1991) Low-Speed Aerodynamics. New York, NY: McGraw-Hill.
[23] Cottet G-H and Koumoutsakos PD (2000) Vortex Methods: Theory and Practice. Cambridge, UK: Cambridge University Press.
[24] Sørensen JN and Myken A (1992) Unsteady actuator disc model for horizontal axis wind turbines. Journal of Wind Engineering and Industrial Aerodynamics
39: 139–149.
[25] Sørensen JN and Kock CW (1995) A model for unsteady rotor aerodynamics. Journal of Wind Engineering and Industrial Aerodynamics 58: 259–275.
240 Aerodynamic Analysis of Wind Turbines

[26] Vermeer LJ, Sørensen JN, and Crespo A (2003) Wind turbine wake aerodynamics. Program Aerospace Science 39: 467–510.
[27] Hansen MOL, Sørensen JN, Voutsinas S, et al. (2006) State of the art in wind turbine aerodynamics and aeroelasticity. Program Aerospace Science 42: 285–330.
[28] Brouckaert, J-F (ed.) (2007) Wind Turbine Aerodynamics: A State-of-the-Art. VKI Lecture Series 2007-05. Belgium: von Karman Institute for Fluid Dynamics.
[29] Sørensen JN and Shen WZ (2002) Numerical modelling of wind turbine wakes. Journal of Fluids Engineering 124(2): 393–399.
[30] Leclerc C and Masson C (2004) Towards blade-tip vortex simulation with an actuator-lifting surface model. AIAA-2004-0667, American Institute of Aeronautics and Astronautics,
USA.
[31] Mikkelsen R (2003) Actuator Disc Methods Applied to Wind Turbines. PhD Dissertation. DTU Mechanical Engineering.
[32] Troldborg N (2008) Actuator Line Modelling of Wind Turbine Wakes. PhD Dissertation. DTU Mechanical Engineering.
[33] Ivanell SA (2009) Numerical Computations of Wind Turbine Wakes. PhD Dissertation. KTH, Royal Institute of Technology.
[34] Spalart P and Allmaras S (1994) A one-equation turbulence model for aerodynamic flows. La Recherches Aerospace 1(1): 5–21.
[35] Jones WP and Launder BE (1972) The prediction of laminarization with a two-equation model of turbulence. International Journal of Heat and Mass Transfer 15: 301–304.
[36] Menter FR (1993) Zonal two-equation k–ω models for aerodynamic flows. AIAA Paper 93-2906, American Institute of Aeronautics and Astronautics, USA.
[37] Lesieur M, Metais O, and Comte P (2005) Large-Eddy Simulations of Turbulence. Cambridge, UK: Cambridge University Press.
[38] Spalart PR, Jou W-H, Stretlets M, and Allmaras SR (1997) Comments on the feasibility of LES for wings and on the hybrid RANS/LES approach, advances in DNS/LES.
Proceedings of the First AFOSR International Conference on DNS/LES.
[39] Sørensen, JN (ed.) (1999) VISCWIND: Viscous effects on wind turbine blades. Report ET-AFM-9902. Lyngby, Denmark: Department of Energy Engineering, DTU.
[40] Chaviaropoulos PK, Nikolaou IG, Aggelis K, et al. (2001) Viscous and aeroelastic effects on wind turbine blades: The VISCEL project. Proceeding of 2001 European Wind Energy
Conference and Exhibition. Copenhagen, 2–6 July.
[41] Sørensen NN and Hansen MOL (1998) Rotor performance predictions using a Navier-Stokes method. AIAA Paper 98-0025, American Institute of Aeronautics and Astronautics,
USA.
[42] Duque EPN, van Dam CP, and Hughes S (1999) Navier-Stokes simulations of the NREL combined experiment phase II rotor. AIAA Paper 99-0037, American Institute of
Aeronautics and Astronautics, USA.
[43] Sørensen NN, Michelsen JA, and Schreck S (2002) Navier-Stokes predictions of the NREL phase VI rotor in the NASA-AMES 80 ft  120 ft wind tunnel.
Wind Energy 5: 151–169.
[44] Schreck S (2002) The NREL full-scale wind tunnel experiment introduction to the special issue. Wind Energy 5(2–3): 77–84.
[45] Sørensen NN (2009) CFD modelling of laminar-turbulent transition for airfoils and rotors using the γ–Reθ model. Wind Energy 12(8): 715–733.
[46] Sørensen JN (2011) Aerodynamic aspects of wind energy conversion. Annual Review of Fluid Mechanics 43: 427–448.
[47] Sørensen, JN, Hansen, MOL and Hansen, KS (eds.) (2007) The science of making torque from wind. Journal of Physics: Conference Series 75.
[48] Ainslie JF (1985) Development of an eddy viscosity model for wind turbine wakes. Proceeding of 7th BWEA Wind Energy Conference. Oxford, 27–29 March.
[49] Crespo A and Hernández J (1996) Turbulence characteristics in wind-turbine wakes. Journal of Wind Engineering and Industrial Aerodynamics 61(1): 71–85.
[50] Ivanell S, Sørensen JN, Mikkelsen R, and Henningson D (2008) Analysis of numerically generated wake structures. Wind Energy 12(1): 63–80.
[51] Ivanell S, Mikkelsen R, Sørensen JN, and Henningson D (2010) Stability of the tip vortices of a wind turbine. Wind Energy 13(8): 705–715.
[52] Troldborg N, Larsen GL, Madsen HA, et al. (2010) Numerical simulations of wake interaction between two wind turbines at various inflow conditions. Wind Energy DOI: 10.1002/
we.433.
[53] Troldborg N, Sørensen JN, and Mikkelsen R (2009) Numerical simulations of wake characteristics of a wind turbine in uniform flow. Wind Energy 13(1): 86–99.
[54] Crespo A, Hernandez J, and Frandsen S (1998) A survey of modelling methods for wind-turbine wakes and wind farms. Wind Energy 2: 1–24.
[55] Theodorsen T (1948) Theory of Propellers. New York, NY: McGraw-Hill Book Company.
[56] Okulov VL and Sørensen JN (2008) Refined Betz limit for rotors with a finite number of blades. Wind Energy 11: 415–426.
[57] Okulov VL and Sørensen JN (2010) Maximum efficiency of wind turbine rotors using Joukowsky and Betz approaches. Journal of Fluid Mechanics 649: 497–508.
[58] Fuglsang P and Madsen HA (1999) Optimization method for wind turbine rotors. Journal of Wind Engineering and Industrial Aerodynamics 80: 191–206.
[59] Fuglsang P and Thomsen K (2001) Site-specific design optimization of 1.5–2.0 MW wind turbines. Journal of Solar Energy Engineering 123: 296–303.
[60] Xudong W, Shen WZ, Zhu WJ, et al. (2009) Shape optimization of wind turbine blades. Wind Energy 12(8): 781–803.
[61] Wagner S, Bareiss R, and Guidati G (1996) Wind turbine noise. EUR 16823. Berlin, Germany: Springer.
[62] Brooks TF, Pope DS, and Marcolini MA (1989) Airfoil self-noise and prediction. NASA Reference Publication 1218. USA: National Aeronautics and Space
Administration.
[63] Amiet RK (1975) Acoustic radiation from an airfoil in a turbulent stream. Journal of Sound and Vibration 41: 407–420.
[64] Fuglsang P and Madsen HA (1996) Implementation and verification of an aeroacoustic noise prediction model for wind turbines. Risø National Laboratory Publication R-867(EN),
Risø National Laboratory, Denmark.
[65] Zhu WJ, Heilskov N, Shen WZ, and Sørensen JN (2005) Modeling of aerodynamically generated noise from wind turbines. Journal of Solar Energy Engineering
127: 517–528.
[66] Zhu WJ, Sørensen JN, and Shen WZ (2005) An aerodynamic noise propagation model for wind turbines. Wind Engineering 29(2): 129–143.
[67] Schepers JG, Curvers APWM, Oerlemans S, et al. (2005) SIROCCO: Silent Rotors by Acoustic Optimisation. First International Meeting on Wind Turbine Noise: Perspectives for
Control.
[68] Moriarty P, Guidati G, and Migliore P (2005) Prediction of turbulent inflow and trailing-edge noise for wind turbines. AIAA Paper 2005–2881, Proceedings of the 11th AIAA/CEAS
Aeroacoustics Conference. Monterey, CA.
[69] Lutz T, Herrig A, Würz W, et al. (2007) Wind-tunnel verification of low-noise airfoils for wind turbines. AIAA Journal 45(4): 779–785.
[70] Drela M (1989) XFOIL: An analysis and design system for low Reynolds number airfoils. Conference on Low Reynolds Number Aerodynamics. University Notre Dame.

Further Reading

[1] Burton T, Sharpe D, Jenkins N, and Bossanyi E (2001) Wind Energy Handbook. New York, NY: Wiley.
[2] Hansen AC and Butterfield CP (1993) Aerodynamics of horizontal-axis wind turbines. Annual Review of Fluid Mechanics 25: 115–149.
[3] Hansen MOL (2008) Aerodynamics of Wind Turbine. London, UK: Earthscan.
[4] Hansen MOL, Sørensen JN, Voutsinas S, et al. (2006) State of the art in wind turbine aerodynamics and aeroelasticity. Progress in Aerospace Sciences 42: 285–330.
[5] Hau E and von Renouard H (2006) Wind Turbines: Fundamentals, Application, Economics. Berlin, Germany: Springer.
[6] Leishman JG (2002) Challenges in modeling the unsteady aerodynamics of wind turbines. Wind Energy 5: 86–132.
[7] Manwell F, McGowan JG, and Rogers AL (2010) Wind Energy Explained: Theory, Design and Application. West Sussex, UK: Wiley.
[8] Snel H (1998) Review of the present status of rotor aerodynamics. Wind Energy 1: 46–69.
Aerodynamic Analysis of Wind Turbines 241

[9] Sørensen, JN and Sørensen, JD (eds.) (2011) Wind Energy Systems: Optimising Design and Construction for Safe and Reliable Operation. Cambridge, UK: Woodhead Publishing
Series in Energy No. 10.
[10] Sørensen JN (2011) Aerodynamic aspects of wind energy conversion. Annual Review of Fluid Mechanics 43: 427–448.
[11] Spera DA (1994) Wind Turbine Technology. New York, NY: ASME Press.
[12] Vermeer LJ, Sørensen JN, and Crespo A (2003) Wind turbine wake aerodynamics. Progress in Aerospace Sciences 39: 467–510.
[13] De Vries O (1979) Fluid dynamic aspects of wind energy conversion. AGARD Report AG-243, AGARD Advisory Group for Aerospace Research & Development.

You might also like