Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

2.

07 Aerodynamic Analysis of Wind Turbines


Jens N Sørensen, DTU Wind Energy, Section of Aero and Fluid Dynamics, Lyngby, Denmark
r 2022 Elsevier Ltd. All rights reserved.

2.07.1 Introduction 172


2.07.2 Momentum theory 173
2.07.2.1 One-dimensional momentum theory 173
2.07.2.2 The optimum rotor of Glauert 174
2.07.2.3 The blade-element momentum theory 175
2.07.2.3.1 Tip correction 176
2.07.2.3.2 Correction for heavily loaded rotors 177
2.07.2.3.3 Yaw correction 177
2.07.2.3.4 Dynamic wake 177
2.07.2.3.5 Airfoil data 177
2.07.3 Optimum Glauert model for finite-bladed rotor 178
2.07.4 Design of Glauert rotor 179
2.07.5 Advanced aerodynamic modeling 180
2.07.5.1 Vortex models 180
2.07.5.2 Numerical actuator disk and line models 181
2.07.5.3 Full Navier-stokes modeling 182
2.07.6 CFD computations of wind turbine rotors 183
2.07.7 CFD in wakes and wind farms 185
2.07.8 Rotor optimization using BEM technique 187
2.07.9 Noise from wind turbines 189
References 191
Further reading 193

2.07.1 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 aero-elastic model for predicting performance and structural deflections is a
prerequisite for the 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 (Lanchester, 1915) and Betz (Betz, 1920) predicted
that even an ideal wind turbine cannot exploit more than 59.3% of the wind power passing through the rotor disk. A major break-
through in rotor aerodynamics was achieved by Betz (Betz, 1920) and Glauert (Glauert, 1935), 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 the industry.
From an outsider's point of view, aerodynamics of wind turbines may seem simple as compared to aerodynamics of e.g. 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 it, leading to flow patterns far more
complex than that of non-separated 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 be determined within an acceptable accuracy. When boundary layer separation occurs, the centrifugal force

172 Comprehensive Renewable Energy, Second Edition, Volume 2 doi:10.1016/B978-0-12-819727-1.00127-8


Aerodynamic Analysis of Wind Turbines 173

Uo uR u1

Fig. 1 Control volume for one-dimensional actuator disk.

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
non-rotating 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
3-dimensional and unsteady flow behavior.
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 the 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 forms the backbone in rotor design of wind turbines, is introduced. Next, state of the art advanced
aerodynamic models are presented. These include 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.

2.07.2 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 a brief introduction to momentum theory for
the design and analysis of wind turbines we will given, starting by the simple, albeit important, one-dimensional momentum
theory, from which the Betz limit can be derived, and ending with the practical Blade Element Momentum (BEM) theory, which
forms the basis for all rotor design codes in use by the industry.

2.07.2.1 One-dimensional momentum theory


Firstly, the simple axial momentum theory is revisited as it originated by Rankine (Rankine, 1865), Froude (Froude, 1889) and
Froude (Froude, 1878). Consider an axial flow of speed Uo that 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 ¼ po, and let r denote the density of air. We now consider a one-dimensional
model for the stream tube that encloses the rotor disk (see Fig. 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,m, _ is constant in each cross-section. Thus,

_ ¼ rUo Ao ¼ ruR A ¼ ru1 A1 :


m ð1Þ

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

_ ðUo  u1 Þ ¼ ruR AðUo  u1 Þ:


T¼m ð2Þ

Applying the Bernoulli equation in front of and behind the rotor, the total pressure head of the air in the slipstream has been
decreased by
 
Dp ¼ ½r Uo2  u21 : ð3Þ

The pressure drop takes place across the rotor and represents the thrust, T ¼ ADp. Combining Eqs. (2), (3) shows the well-
known result that

uR ¼ ½ðu1 þ Uo Þ: ð4Þ
174 Aerodynamic Analysis of Wind Turbines

Introducing the axial interference factor as follows


Uo  uR
a¼ ; ð5Þ
Uo
it results that uR ¼ (1  a)Uo and u1 ¼ (1  2a)Uo. From Eq. (2) one gets the following expressions for thrust and power
extraction
T ¼ 2rAU2o að1  aÞ; ð6Þ

P ¼ uR T ¼ 2rAUo3 að1  aÞ2 : ð7Þ


Introducing the dimensionless thrust and power coefficient, respectively,
T P
CT  ; CP  ; ð8Þ
½rAU 2o ½rAU3o
it results
CT ¼ 4að1  aÞ; CP ¼ 4að1  aÞ2 : ð9Þ
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 : ð10Þ
27 3
This result is usually referred to as the Betz limit or the ‘Lanchester-Betz-Joukowsky limit,’ as recently proposed by van Kuik
(van Kuik, 2007), 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.

2.07.2.2 The optimum rotor of Glauert


Utilizing general momentum theory Glauert (Glauert, 1935), 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 sizeDr, Eq. (2) then reads,
DT ¼ 4prUo2 ð1  aÞarDr: ð11Þ
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,
Dp ¼ rOruy þ ½ru2y : ð12Þ
where O is the angular velocity.
Combining Eqs. (11), (12), one gets
DT ¼ Dp  2prDr ¼ 2pruy ðOr þ ½uy ÞrDr; ð13Þ
where uy is the azimuthal velocity behind the rotor. Defining the azimuthal interference factor as,
0 uy
a ¼ ; ð14Þ
2Or
Eq. (13) reads,
 0
 0
DT ¼ 4prO2 1 þ a a r 3 Dr: ð15Þ

Combining Eqs. (11), (15), it results in:


 0
 0
ð1  aÞa ¼ l2 x2 1 þ a a ; ð16Þ

where x ¼ r/R and l ¼ OR/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, the following
expression for the useful power produced by the wind turbine is derived:
ZR Z1
0
P¼O 2pr ruuy dr ¼ 4prO Uo
2 2
a ð1  aÞx3 dx; ð17Þ
0 0

or in dimensionless form,
Z 1
0
CP ¼ 8l2 a ð1  aÞx3 dx: ð18Þ
0
Aerodynamic Analysis of Wind Turbines 175

Table 1 Flow conditions for the optimum actuator disk.

a a0 a0 x2l2 lx

0.25 1 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 0.00301 0.2216 8.58
1/3 0 0.2222 1

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, 1935, or Wilson and Lissaman, 1974). This results in the following relation for an
optimum rotor,
0
da 0
ð1  aÞ  a ¼ 0: ð19Þ
da
Differentiating Eq. (16) with respect to a gives,
 0
 da0
1  2a ¼ l2 x2 1 þ 2a : ð20Þ
da
Combining Eqs. (16), (19), (20) results in the following expression:
0 1  3a
a ¼ : ð21Þ
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 r a r 1/3.
The relations between a, a0 , a0 x2l2 and lx for an optimum rotor are given in Table 1. The maximal power coefficient as a
function of the tip speed ratio is determined by integrating Eq. (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.

2.07.2.3 The blade-element momentum theory


The Blade Element Momentum (BEM) method was developed by Glauert (Glauert, 1935) as a practical way to analyse and design
rotor blades. In the BEM theory the loading is computed using two independent methods, i.e. by a local blade element con-
sideration using tabulated two-dimensional (2D) airfoil data and by the use of the one-dimensional momentum theorem. First,
employing blade-element theory, axial load and torque are written as, respectively,
dT 1
¼ BF n ¼ rcBV 2rel  Cn ; ð22Þ
dr 2
dM 1
¼ BrF t ¼ rcBrV 2rel  Ct ; ð23Þ
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 2-dimensional tabulated force coefficients.
From the velocity triangle at the blade-element (see Fig. 2), it can be deduced that
 0
U1 ð1  aÞ Or 1 þ a
sin f ¼ ; cos f ¼ ; ð24Þ
Vrel Vrel
where the induced velocity is defined as Wi ¼ (  aU0,a0 Or). Using the above relations, one gets
 0
U 2 ð1  aÞ2 U1 ð1  aÞOr 1 þ a
2
Vrel ¼ 1 2 ¼ : ð25Þ
sin f sin fcos f
Inserting these expressions into Eqs. (22), (23), it results in:
dT rBcU 20 ð1  aÞ2
¼  Cn ; ð26Þ
dr 2sin2 f
176 Aerodynamic Analysis of Wind Turbines

Table 2 Power coefficient as a function of the tip speed ratio for an optimum
actuator disk.

l CPmax

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

Fig. 2 Cross sectional airfoil element.


 0
dM rBcU 0 ð1  aÞOr 2 1 þ a
¼  Ct : ð27Þ
dr 2sin fcos f
Next, applying axial momentum theory, the axial load is computed as
dT
¼ rðU0  uwake Þ2pruR ¼ 4prU02 að1  aÞ; ð28Þ
dr
where uR ¼ U0(1  a) is the axial velocity in the rotor plane and uwake ¼ U0(1  2a) is the axial velocity in the ultimate wake.
Applying the moment of momentum theorem, it results in:
dM 0
¼ rruy 2pruR ¼ 4prr 3 OU0 a ð1  aÞ; ð29Þ
dr
where uy ¼ 2Ora0 is the induced tangential velocity in the far wake. Combining Eqs. (26), (27) with Eqs. (28), (29), one gets after
some algebra
1
a¼ ; ð30Þ
4sin2 f=ðsCn Þ þ 1

0 1
a ¼ ; ð31Þ
4sin fcos f=ðsCt Þ  1
where s ¼ Bc/2pr is the rotor solidity at radius “r”.

2.07.2.3.1 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, Glauert3 introduced Prandtl's tip loss factor. In this method a
correction factor, F, is introduced which corrects the loading. In a recent paper by Shen et al. (Shen et al., 2005) the tip correction is
discussed and various alternative formulations are compared. However, here the correction is limited to the original form given by
Glauert (Glauert, 1935). In this model the induced velocities are corrected by the tip loss factorF, modifying Eqs. (28), (29) as
follows,
dT
¼ 4prrU1
2
aF ð1  aÞ ð32Þ
dr
dM 0
¼ 4prr 3 OU1 a F ð1  aÞ; ð33Þ
dr
Aerodynamic Analysis of Wind Turbines 177

An approximate formula of the Prandtl tip loss function was introduced as follows,
  
2 BðR  r Þ
F ¼ cos1 exp  ; ð34Þ
p 2rsin f
where f ¼ f(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 ¼ Cl cos f þ Cd sin f and Ct ¼ Cl sin f
 Cd cos f, respectively. (Cl, Cd) depend on local airfoil shape and are obtained using tabulated 2D airfoil data corrected with 3D
rotating effects. Equating Eq. (26) to Eq. (32), and Eq. (27) to Eq. (33), the final expressions for the interference factors read
1
a¼ ; ð35Þ
4F sin2 f=ðsCn Þ þ 1
0 1
a ¼ ; ð36Þ
4F sin fcos f=ðsCt Þ  1

2.07.2.3.2 Correction for heavily loaded rotors


By putting Eq. (32) into dimensionless form the following expression for the local thrust coefficient is derived:
dT
CT ¼ ¼ 4aF ð1  aÞ: ð37Þ
½rU1
2 2prdr

For heavily loaded rotors, i.e. 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
Glauert3, an appropriate correction is to replace the expression for a Z 1/3 with the following expression:
 a 
CT ¼ 4aF 1  ð5  3aÞ : ð38Þ
4
As discussed in e.g. Spera (Spera, 1994) or Hansen (Hansen, 2008) other expressions can also be used.

2.07.2.3.3 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 (Snel and Schepers, 1995),
 r χ  
wi ¼ wi0 1 þ tan cos ðyblade  y0 Þ ; ð39Þ
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 yblade denotes the azimuthal
position of the blade and y0 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 (Hansen, 2008).

2.07.2.3.4 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 (Schepers and Snel, 1995)) 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,
dui DT
Rf ðr=RÞ þ 4ui ðU0  ui Þ ¼ ; ð40Þ
dt 2prDr
where the function f(r/R) is a semi-empirical function associated with the induction. The equation can be seen to correspond to the
axial momentum equation, Eq. (28), except for the time term which is responsible for the time delay.

2.07.2.3.5 Airfoil data


As a pre-step to the BEM computations, 2-dimensional 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 three-
dimensional and rotational effects. A simple correction formula for rotational effects was proposed by Snel and van Holten (Snel
and van Holten, 1995) for incidences up to stall. For higher incidences (4401), 2-dimensional lift and drag coefficients of a flat
178 Aerodynamic Analysis of Wind Turbines

Axial interference factor with and without tip correction


0.4
TSR=1
TSR=3
TSR=6
TSR=10
0.35
a [–]

0.3

0.25
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
x [–]
Fig. 3 Axial interference factor as a function of radial distance for different tip speed ratios. Solid lines: with tip correction; Dashed lines: without
tip correction.

plate can be used. These data, however, are too big because of aspect ratio effects and here the correction formulas of Viterna and
Corrigan (Viterna and Corrigan, 1981) are usually applied (see also Spera (Spera, 1994)). 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 (Øye, 1991), or it can be considerable more advanced,
taking into account also the attached flow, leading edge separation and compressibility effects, as in the model of Leishman and
Beddoes (Leishman and Beddoes, 1989).

2.07.3 Optimum Glauert model for finite-bladed rotor

Following the principles outlined in Section 2.07.2.2 for a rotor disk with infinitely many blades, a more correct model to include
a finite number of blades can be derived using the tip correction. From Eq. (33), the power coefficient may be derived as
Z 1
0
CP ¼ 8l2 a F ð1  aÞx3 dx; ð41Þ
0

where F is the tip correction, shown in Eq. (34), and x ¼ r/R is the dimensionless radial coordinate. Like in Section 2.07.2.2 the
optimum rotor condition may be determined by determining the optimum of the integrand, H ¼ a0 F(1  a), with the constraint
given from Eq. (16). This constitutes two equations with two unknowns, a and a0 . However, due to the complex nature of the tip
correction formula, it is not possible to derive a simple analytical solution, like in Eq. (21). Instead, we resort to a numerical
solution. Hence, the problem to be solved is stated as follows:
For fixed tip speed ratio, l, determine the values of a and a0 for which H ¼ a0 F(1  a) attains a maximum and (1  a)
a ¼ l2x2(1 þ a0 )a0 .
The solution to the problem is shown graphically in Fig. 3, where the optimum interference factor a is shown as a function of
the radial distance for different tip speed ratios, and compared to the value obtained without tip correction (see also Table 1). As
expected, the values are nearly identical far from the tip, where the tip correction is unity, whereas large differences are observed
near the tip. It is interesting that the curves at the tip tend to a value a ¼ 0.4, whereas the values without tip correction tends
toward 1/3 (see Table 1 and Eq. 21). To explain this behavior, a simple limit analysis may be performed. Assuming that l - 1 and
x - 1, Eqs. (16), (24) may be approximated as

0 ð1  aÞa 1a
aE and sin fE ; ð42Þ
l2 x 2 lx

which, inserted into Eq. (34) gives


  
2 Nb l
FE cos1 exp  ð1  xÞ : ð43Þ
p 2ð1  aÞ
pffiffiffiffiffiffiffi Nb l
Exploiting that exp(  cy) ¼ 1  cy þ o(cy)2 and cos1 ð1  cyÞ ¼ 2cy þ oðcyÞ3=2 , with c ¼ 2ð1aÞ and y ¼ 1  x, Eq. (43) is
approximated as
Aerodynamic Analysis of Wind Turbines 179

Fig. 4 Dimensionless rotor geometry of optimum Glauert rotor with and without tip correction. Solid lines: with tip correction; Dashed lines:
without tip correction.
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2 Nb l
FE ð1  xÞ: ð44Þ
p ð1  aÞ

Hence, it results that


sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
0 2 Nb l að1  aÞ
H ¼ a F ð1  aÞE ð1  xÞ  : ð45Þ
p ð1  aÞ l2 x2
The optimum power coefficient is obtained by differentiating H with respect to a, and setting the result equal to zero,
∂H ∂ h i 3
¼0 ) að1  aÞ3=2 ¼ 0 ) ð1  aÞ3=2  að1  aÞ1=2 ¼ 0 ) a ¼ 0:4;
∂a ∂a 2
which actually verifies the limits of the axial interference factor displayed in Fig. 3.

2.07.4 Design of Glauert rotor

As described before, an optimum rotor is designed by determining the combination of induced velocities that maximizes the
power coefficient. Having determined the induced velocities, the planform and twist can be determined for a given tip speed ratio.
The fundamental expressions for the forces acting on a blade (Fig. 2) are most conveniently expressed by the Kutta–Joukowsky
theorem,
L ¼ rVrel G; ð46Þ
where L is the lift force on a blade element and r is the density of the air.
Introducing the local lift coefficient,
L
Cl ¼ ð47Þ
2 rcV rel ;
1 2

where c is the local chord length, one gets


G ¼ 1 =2 Vrel cCl : ð48Þ
Combining Eqs. (38), (48), gives the formula:
2prFuy ¼ 1 =2 Vrel Nb cCl ; ð49Þ
from which the chord distribution can be determined
0
c 8plxFa
¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ; ð50Þ
R
Nb Cl ð1  aÞ2 þ l2 x2 ð1  a0 Þ2

where l ¼ OR/U0is the tip speed ratio.


In order not to let the basic design be dependent on the lift coefficient, it is convenient to re-arrange this equation as
0
4lx2 Fa
sCl ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ; ð51Þ
2
ð1  aÞ2 þ l2 x2 ð1  a0 Þ
180 Aerodynamic Analysis of Wind Turbines

Fig. 5 Design flow angle for optimum 3-bladed Glauert rotor as a function of radial distance.

where s ¼ Nbc/(2pr) is the local solidity. The dimensionless optimum plan form for different design tip speed ratios are in Fig. 4
shown for designs both with and without tip correction. As a typical example, the rotor is 3-bladed and designed at a unity lift
coefficient. From the figure it is clearly seen that the main differences in the plan form appear for small tip speed ratios, where the
tip correction is pronounced over most of the rotor area.
In a practical design, the lift coefficient is typically kept constant at all cross sections and taken as a compromise between having
a maximum lift and an optimum lift/drag ratio. At high Reynolds numbers the contribution of the drag does not change the
design, hence Eq. (51) is sufficient for designing the plan form of an optimum rotor for a given tip speed ratio. From the lift
coefficient a design angle of attack, aD, is derived from which the design twist angle is determined as
b ¼ f  aD ; ð52Þ
where the flow angle, f, is determined from (see Fig. 2)
 
1a
f ¼ tan1 0 : ð53Þ
lxð1 þ a Þ
In Fig. 5 the design flow angle distribution for a 3-bladed rotor is depicted. From the figure it is seen that the tip correction only
has a slight influence on the design flow angle.

2.07.5 Advanced aerodynamic modeling

Although the BEM method is widely used and today constitutes the only design methodology in use by the 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.

2.07.5.1 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 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 (Joukowsky, 1912), 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 (Margoulis, 1922)). 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 two-
dimensional airfoil data. This can be accomplished by representing the span wise loading on each blade by a series of straight
Aerodynamic Analysis of Wind Turbines 181

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 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 semi-empirical rules. This
makes the technique fast to be used 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 wake vortex methods tend
to suffer from stability problems owing to the intrinsic singularity in induced velocities that appears when vortex elements are
approaching each other. 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 (Leishman, 2002)). A special version of the free vortex wake methods is the
method by Voutsinas (Voutsinas, 2006) in which the flow modeling is taken care of by vortex particles or vortex blobs.
A generalization of the vortex method is the so-called Boundary Integral Equation 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 (Katz and Plotkin, 1991)) or Cottet and Koumoutsakos (Cottet and Koumoutsakos, 2000).
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.
A new development in viscous-inviscid based panel methods for wind turbine computations were introduced by Ramos-Garcia
et al. (Ramos-García et al., 2014b, 2016, 2017). In this model, a new approach that couples a 3D panel method and a free-wake
vortex model with a boundary layer solver was developed. The solver is based on a viscous-inviscid coupling technique that is
performed locally at each spanwise cross section by using the transpiration velocity concept. The viscous part is based on the quasi
3-dimensional boundary layer solver, Q3UIC (Ramos-Garcia et al., 2014a), which is a strongly coupled viscous-inviscid technique
that includes rotational effects in the boundary layer. The wake is modeled using straight vortex lines to resolve the wake dynamics.
A viscous core and filament straining model was included to take into account the viscous effects in the wake. The coupling
between the viscous and the inviscid parts is carried out locally along the blade span using the transpiration velocity method,
which introduces a new source distribution at the body surface that forces the outwards displacement of the limiting streamlines.
In this way, the effect of viscosity and rotation inside the boundary layer of the rotor blade is taken into account by modifying the
potential flow solution.
Today, various aerodynamic simulation and design tools exists as open software, including lifting line models for wind
turbines. An example of this is the QBlade software developed at TU Berlin (Marten et al., 2016). Besides a basic BEM solver and
modules for analysis and design of 2D airfoils, the code also contains a lifting line model for more sophisticated analyses of rotor
aerodynamics (Marten et al., 2013).

2.07.5.2 Numerical actuator disk and line 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 one-
dimensional momentum theory. Combining it with a blade-element analysis, it results in 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
182 Aerodynamic Analysis of Wind Turbines

in the sense that the aerodynamic forces have to be determined from measured airfoil characteristics, corrected for three-
dimensional 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 (Sørensen and Myken, 1992) and Sørensen and Kock (Sørensen and Kock, 1995). 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. (Vermeer et al., 2003), Hansen et al. (Hansen et al., 2006) or the VKI Lecture Series (Brouckaert, 2007).
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 three-
dimensional actuator disk model has been developed Sørensen and Shen (Sørensen and Shen, 2002). The model combines a three-
dimensional 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 three-dimensional 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
(Leclerc and Masson, 2004). A main motivation for developing such types of model is to be able to analyse and verify the validity of
the basic assumptions that are employed in the simpler more practical engineering models and to perform wind farm simulations
without having to resolve the rotor boundary layer. Reviews of the basic modeling of actuator disk and actuator line models can be
found in the PhD dissertations of Mikkelsen (Mikkelsen, 2003), Troldborg (Troldborg, 2008) and Ivanell (Ivanell, 2009).
In the past 15 years, the actuator line method has evolved from a specialized academic tool into a standard technique for high-
fidelity computations of full wind farms subject to atmospheric boundary layer flows. In its early development, there were several
shortcomings associated with the method. In particular, the use of a regularization function to smear out the loading caused
problems of representing adequately the flow near the tip. In the original formulation of the model, the smearing function was
introduced through a three-dimensional Gaussian to avoid numerical instabilities. In later developments, a more physical
explanation of the force projection function has been related to the geometry of the blade. Shives and Crawford (Shives and
Crawford, 2013) proposed a model in which the isotropic Gaussian varies in radial direction, with the Gaussian width being
proportional to the airfoil chord length. Based on this, to reduce the over-predicted forces near the tip, Jha et al. (Jha et al., 2014)
introduced an equivalent elliptic planform chord as a measure to determine the Gaussian width. To come up with a more physical
smearing function, Churchfield et al. (Churchfield et al., 2017) introduced an anisotropic Gaussian force projection, which width
is proportional to the airfoil sectional geometry for both chord length and airfoil thickness. Their results showed that the
anisotropic force projection generates a more realistic and compact blade wake when compared to a constant projection width
along the blade. Using a 2D approach, Martinez-Tossas et al. (Martinez-Tossas et al., 2017). showed that the optimum width for an
anisotropic force projection lies in the range of 0.2–0.4 chord lengths, as compared to an analytical airfoil solution. Jha et al. (Jha
and Schmitz, 2018) modified a 2D kernel function to be applied to curved blade geometries.
The effort to achieve higher fidelity of the actuator line method through a more sophisticated smearing functions, however, has
the price that it demands more node points to represent the rotor blade, meaning that the computational time increases. Another
way of tackling with the problem is to look for the cause of the overestimated loads near the tip. In the recent work of Dag (Dag,
2017) and Dag and Sørensen (Dag and Sørensen, 2020) it was shown that it is the resulting thickness of the trailing vortices from
the projection method that causes problems. The problem was solved by introducing a correction to the local angle of attack to
account for the over-corrected induction from the projection method. In a paper by Forsting et al. (Forsting et al., 2020), an
improved formulation of the correction technique was proposed and successfully validated. A theoretical explanation of the
concept has been offered by Martinez-Tossas and Meneveau (Martinez-Tossas and Meneveau, 2019). Another shortcoming of the
actuator line method, which is general for lifting line methods, is that its accuracy degenerates for lifting surfaces at low aspect
ratios. This can to a certain degree be counteracted by using smearing functions that take into account the actual geometry of the
surface. However, an alternative method in line with the correction proposed by Dag (Dag, 2017) and Dag and Sørensen (Dag and
Sørensen, 2020) for the over-estimated loads, is to correct for the aspect ratio through the influence of the trailing vortices. Such a
method has been introduced by Sørensen et al. (Sørensen et al., 2016) using a so-called decambering correction.

2.07.5.3 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
Computational Fluid Dynamics (CFD) tools based on the solution of the Navier-Stokes equations. Within aerodynamics this
research has mostly been related to flows around fixed-wing aircrafts 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, 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.
Aerodynamic Analysis of Wind Turbines 183

Table 3 Number of required mesh points for various types of computations.

Airfoil Full Rotor Wake

RANS 10 5
107
105  108
DES 107 108 107  1010
LES 1010 1012 107  1014
DNS 1015 1017 107  1019

Some of the experience gained from the aeronautical research institutions has been exploited directly in the development of
CFD algorithms for wind turbines. Notable 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 for a long time have existed 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 (Spalart and Allmaras, 1994) or the k   (Jones and Launder, 1972) model. By using this kind of
model only the time-averaged flow field is 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 3-
D 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  o SST model developed by Menter (Menter, 1993) 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 models lack of ability of representing a full unsteady
spectrum. Thus, for attached flow the accuracy is fully adequate whereas it for stalled flows 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 Large Eddy Simulation (LES). In LES the Navier-Stokes equations are filtered
spatially on the computational mesh and only the sub-grid scale part of the turbulence is modeled using a so-called Sub-Grid-Scale
(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 the large number of mesh points 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 Lesieur et al. (Lesieur et al., 2005) an estimate of the ratio between the largest and the smallest length scale can be
given as L/ℓ E ReL3/4, where the Reynolds number ReL ¼ UL/u, with U denoting a characteristic wind speed and u is the kinematic
viscosity. For an airfoil of a wind turbine blade a typical value is ReC D 5  106 where index c denotes the chord length. Thus, for a
DNS computation of an airfoil section in the order of 105 mesh points are required 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, assuming
that the sub-grid scale covers about 1.5 decades. A main difference between RANS and LES is that RANS computations may be
carried out in a pure 2-D domain, e.g. when studying or designing airfoils, whereas LES is always intrinsically unsteady and 3-D. As
a compromise between the fast computing time of RANS methods and the accuracy of LES, Spalart et al. (Spalart et al., 1997)
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 to be depended on the
Reynolds number, if e.g. 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 a 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 D 107. An overview of the
required number of mesh points for different approaches is given in Table 3.

2.07.6 CFD computations of wind turbine rotors

The research on CFD in wind turbine aerodynamics was in Europe initiated through EU-sponsored collaborate projects, such as
VISCWIND (Sørensen, 1999), VISCEL (Chaviaropoulos et al., 2001) and KNOW-BLADE. The first full Navier-Stokes simulation for
a complete rotor blade was carried out by N.N. Sørensen and Hansen (Sørensen and Hansen, 1998) and later followed by Duque
184 Aerodynamic Analysis of Wind Turbines

Fig. 6 Sketch of limiting streamlines on the suction side of a wind turbine blade of the NREL Phase VI rotor. The cases shown are, from the top,
7, 10 and 20 m/s. The five vertical lines indicate the location of the 30%, 47%, 63%, 80% and 95% sections (Sørensen et al., 2002).

et al. (Duque et al., 1999, 2003), Pape and Lecanu (Pape and Lecanu, 2004) and N.N. Sørensen et al. (Sørensen et al., 2002) in
connection with the American NREL experiment at NASA Ames and the accompanying NREL/NWTC Aerodynamics Blind
Comparison test (Schreck, 2002). 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% (Jha and Schmitz, 2018). 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 modern wind turbine rotors. The Navier-Stokes computations by N.N. Sørensen
et al. (Sørensen, 2009) generally exhibited good agreement with the measurements up to wind speeds of about 10 m/s. 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, 2011). A large number of full 3-D Navier-Stokes computations have later
been carried out by different research groups. The computations include RANS and DNS 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. (Hansen et al., 2006) and Sørensen (Sørensen, 2011), and various contributions were published in the proceedings from
TWIND2007 (Sørensen et al., 2007). To illustrate the degree of complexity one obtains using a full 3-D Navier-Stokes metho-
dology in Fig. 6 it is shown a computation of a rotating 19.1 m long wind turbine blade. It is here clearly seen that a complicated
flow topology results, including a large separated area, which could not be obtained using the BEM technique or inviscid
computations.
With the development of high performance computers, CFD approaches have been employed to gain more insights into the
physics of the flow past rotating blades of wind turbines and to aid in the development of fast engineering type aerodynamics
models. An example of the latter is the understanding and modeling of rotational effects in airfoil data for use in the BEM model.
Using RANS CFD models, airfoil characteristics subject to rotational augmentation have been derived by Johansen and Sørensen
Aerodynamic Analysis of Wind Turbines 185

(Johansen and Sørensen, 2004), Guntur and Sørensen (Guntur and Sørensen, 2015), and Bangga et al. (Bangga et al., 2016, 2017).
Another type of studies concern the understanding and modeling of yaw and shear flows, where properly resolved CFD simu-
lations may aid the development of yaw corrections in the BEM model. Examples of this can be found in Sørensen et al. (Sørensen
et al., 2014) and Schulz et al. (Schulz et al., 2017).
Wind turbines are operating in atmospheric boundary layers where the flow is inherently turbulent and unsteady, hence fluid-
structure coupling is important. However, due to the computational burden of carrying out fully fluid-structure couplings, only
few CFD studies have been conducted up to now. Zahle et al. (Zahle et al., 2009) studied wind turbine rotor-tower interaction
using an incompressible overset grid method. Yu and Kwon (Yu and Kwon, 2014) analysed the 5 MW NREL reference wind
turbine by combining the fluid solver with a structural solver. Li et al. (Li et al., 2015) coupled their fluid solver with a multibody
dynamics structural solver and included atmospheric turbulence in the simulations (Li et al., 2015). Santo et al. (Santo et al.,
2020) represented the rotor by shell elements and run computations in sheared inflow, demonstrating a large influence of the
tower shadow in the computations. Gumma et al. (Guma et al., 2021) performed a comprehensive aero-elastic analysis of wind
turbines operating under turbulent inflow conditions, and showed that fatigue is mainly influenced by the turbulent inflow rather
than the inclusion of flexibility. Recently, Grinderslev et al. (Grinderslev et al., 2021) performed aero-elastic simulations of a
2.3 MW wind turbine rotor operating in different complex atmospheric flows and compared high fidelity fluid–structure inter-
action simulations with aero-elastic simulations using BEM theory. The complexity is further increased when studying floating
wind turbines, where the degrees of freedom also includes the movement of sub-structure. Recent high-fidelity fluid-structure
simulations of floating wind turbines include Tran and Kim (Tran and Kim, 2018) and Calderer et al. (Calderer et al., 2014).

2.07.7 CFD in wakes and wind farms

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 analyse 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 last 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 one rotor
diameter downstream. Here, the presence of the rotor is apparent by the number of blades, blade aerodynamics, including stalled
flow, 3–D 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 focussed on
performance and the physical process of power extraction, while the far wake research is more focussed 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. Semi-analytical far wake models
have been proposed to describe the wake velocity after the initial expansion (e.g. Ainslie (Ainslie, 1985)). Detailed numerical
studies of the far wake have been carried out by Crespo and Hernández (Crespo and Hernández, 1996) 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   turbulence model.
As illustrated in Table 3 prohibitively many mesh points are needed if one wishes to carry out LES or DNS simulations of the
wake in an atmospheric boundary layer. However, employing the actuator line technique and representing the ambient turbulence
and shear 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 (Sørensen and Shen, 2002), Ivanell
et al. (Ivanell et al., 2008, 2010) and Troldborg et al. (Troldborg et al., 2009, 2010) In a recent survey by Vermeer et al. (Vermeer
et al., 2003) 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. (Crespo et al., 1998).
Although nowadays most layout of wind farms is based on simple analytical models, such as those by Jensen (Jensen, 1983),
Larsen (Larsen, 1988) or Frandsen et al. (Frandsen et al., 2006), most development in wind farms aerodynamics has mainly been
focused on the Reynolds-averaged Navier–Stokes equations to study the averaged flow behavior in wind farms (see e.g. the review
by Sanderse et al. (Sanderse et al., 2011)). On the other hand, the fast growth of computational power has in the last decade made
important progress in the development, validation, and application of turbulence-resolving CFD tools, and particularly large-eddy
simulation (LES) for wind energy applications (see e.g. the review by Mehta et al. (Mehta et al., 2014)). In LES, the largest eddies
are resolved and only the smallest scales need to be modeled. This makes LES an appealing alternative for carrying out reliable
186 Aerodynamic Analysis of Wind Turbines

Fig. 7 Vortex structures in the near wake after the Tjæreborg wind turbine (Troldborg et al., 2009).

simulations of large wind farms. LES can provide detailed and accurate information about the dynamics of the wake and the
associated turbulence length scales along with power production and turbine loading. Recent advances in computational power
have made LES accessible for that purpose and the technique has been applied to many wind farm projects to evaluate the effects
of various atmospheric conditions, terrain, and turbine placement.
The first LES computations of an actual operating wind farm was carried out by Ivanell (Ivanell, 2009) who, using an actuator
disk and synthesized airfoil data from the Tjæreborg and a Vestas wind turbines, computed the wind field and the performance of
the Danish Horns Rev. farm. A similar study was subsequently performed by Nilsson et al. (Nilsson et al., 2015) using wind
turbine and production data from the Lillgrund wind farm in Sweden. Wu and Porté-Agel (Wu and Porté-Agel, 2011, 2013) carried
out simulations about a single turbine as well as about a model wind farm using LES and actuator disk and line methodologies.
Yang et al. (Yang et al., 2014) developed a large-eddy simulation framework, dubbed as the Virtual Wind Simulator (VWiS), for
simulating turbulent flow over wind turbines and wind farms in complex terrain. Xie and Archer (Xie and Archer, 2015) studied
the wake generated from a single wind turbine in neutral atmospheric boundary layer conditions using a scale-dependent
Lagrangian dynamic model for the sub-grid scale stress and the actuator line model. Sarlak et al. (Sarlak et al., 2015) performed
LES computations of the flow around a double rotor configuration and compared the results to the blind test experiment of
Pierella et al. (Pierella et al., 2014). Other studies using LES are due to Draper et al. (Draper et al., 2016), who performed a
sensitivity analysis in order to evaluate the influence of the sub-grid scale model and the smearing factor of the projection function
in the actuator disk/line technique. Stevens et al. (Stevens et al., 2018) carried out a detailed comparison between wind tunnel
experiments and LES computations using actuator disk and line models, with and without tower and nacelle. For a more
exhaustive discussion of the various contributions, the reader may consult the review articles by Breton et al. (Breton et al., 2017),
Stevens and Meneveau (Stevens and Meneveau, 2017), Porté-Agel et al. (Porté-Agel et al., 2020).
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. (Troldborg et al., 2010) 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. 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 11 per 3 m.
Fig. 7 shows instantaneous vortex structures in the near wake of a rotor operating at a wind speed of W0 ¼ 6 m/s, corre-
sponding to a tip speed ratio of about 12. It is here seen that the wake flow collapses into small scale turbulence about one
diameter behind the rotor. The actual position depends on the loading and the ambient turbulence level. In the simulation in
Fig. 7 the inflow conditions were pure laminar and the collapse is due to intrinsic instabilities of the flow. Fig. 8 shows the
contours of the instantaneous absolute vorticity in the x/R ¼ 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 bound vorticity of the blades is seen to be shed downstream from the rotor in individual vortex tubes. A closer inspection of
Aerodynamic Analysis of Wind Turbines 187

Fig. 8 Downstream development of wake behind wind turbine. Upper figure: W0 ¼ 6 m/s; Middle Figure: W0 ¼ 10 m/s; Lower figure:
W0 ¼ 14 m/s. In all figures the rotor is located to the left (Troldborg et al., 2009).

the vorticity contours at W0 ¼ 6 m/s 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 ¼ 10 m/s and 14 m/s
distinct tip vortices can be observed about 1.5 rotor radii 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 is observed only 2 rotor radii downstream and 5 rotor radii downstream the entire wake flow
completely breaks up. In the case where W0 ¼ 10 m/s the tip vortices is observed to undergo a Kelvin-Helmholz 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 ¼ 14 m/s, but as expected it takes place even further downstream (approximately 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.

2.07.8 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
188 Aerodynamic Analysis of Wind Turbines

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. The way 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 (Glauert, 1935) and Theodorsen (Theodorsen, 1948) and in revised form by Okulov and Sørensen (Okulov and
Sørensen, 2008, 2010). 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 (Fuglsang
and Madsen, 1999). The objective used in their method was to minimize the cost of energy employing multiple constrains.
Generally, multi-objective 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 farm can also be incorporated in the design process of the wind turbine rotors, Fuglsang and Thomsen
(Fuglsang and Thomsen, 2001). For a more complete overview, the reader is referred to the review papers by Schubel and Crossley
(Schubel and Crossley, 2012), Chehouri et al. (Chehouri et al., 2015) or Khlaifat et al. (Khlaifat et al., 2020).
To illustrate the basics of design optimization of wind turbine rotors, in the following the features of an optimization model
developed at DTU and Chong Qing University (Xudong et al., 2009) for optimizing the geometry of wind turbines to maximize
the energy yield are shown. The method is based on combining an aero-elastic model containing 11 degrees of freedom (DOF)
with the Blade Element Momentum (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 of 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: ð54Þ
i¼1 i¼1

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 (Fuglsang and Madsen, 1999) 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 (Fuglsang and Madsen, 1999). 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, our objective is restricted to
the cost from the rotor. Thus, the objective function is defined as:
Crotor
f ðxÞ ¼ COE ¼ ; ð55Þ
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 ; ð56Þ
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
XN
mi  ci;opt
wrotor ¼ ; ð57Þ
i¼1
M tot :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; 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 Viþ1 k
f ðVi oVoViþ1 Þ ¼ exp   exp  ; ð58Þ
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 ¼ 2, corresponding to the Rayleigh distribution. If a wind turbine operates the full 8760 h per year, its annual energy production
Aerodynamic Analysis of Wind Turbines 189

Fig. 9 Chord and twist angle distributions of the original and the optimized Tjæreborg 2 MW rotor.

is computed as

X
N1  
AEP ¼ ½ P ðViþ1 Þ þ P ðV Þi  f ðVi oVoViþ1 Þ  8760; ð59Þ
i¼1

where P(Vi)is the power at wind speed Vi and N denotes the number of wind speeds considered.
As an example of the optimization model, it is shown here 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, 01 and 12.2%,
respectively, and the upper limits are 3.3 m, 81 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 Fig. 9, in which the chord and twist
distributions of the original and the optimized Tjæreborg rotor are compared.
From Fig. 9A, 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 the chord whereas it decreases significantly in the tip region. The twist angle, Fig. 9B, is slightly smaller than the
original distribution. The performance of the optimized rotor is computed using the aerodynamic/aero-elastic 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. The AEP of the optimized rotor is reduced about 4%,
whereas the cost of the optimized rotor is reduced with about 7.1%. Thus, the cost of energy of the Tjæreborg rotor is reduced
about 3.4%.

2.07.9 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. 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. 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
190 Aerodynamic Analysis of Wind Turbines

respect to public acceptance. This is illustrated by the increasing number of conferences concerning wind turbine noise during the
last years.
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. (Wagner et al., 1996)) for a thorough review of various models). As a compromise
between computing speed and accuracy, the most commonly used models are based on semi-empirical expressions. As basis, most
models employ the experimental results and scaling laws on airfoil self-noise by Brooks et al. (Brooks et al., 1989) together with
the turbulence inflow model proposed by Amiet (Amiet, 1975). This includes e.g. the models of Fuglsang and Madsen (Fuglsang
and Madsen, 1996) and Zhu et al. (Zhu et al., 2005a, b) and the model employed in the SIROCCO project (Schepers et al., 2005),
as well as further developments by Moriarty et al. (Moriarty et al., 2005) and Lutz et al. (Lutz et al., 2007).
In the following some of the features of a typical semi-empirical model are shown, such as those referred 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. (Brooks et al., 1989). 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's 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:
  f ðiÞ 
δ M LDh
SPLi ¼ 10 log i þ Fi ðSt Þ þ Gi ðReÞ; ð60Þ
r2
where δi is the boundary layer displacement thickness, M is the Mach number, f(i) is the raised power which depends on the
particular noise mechanism i, L is the airfoil section semi-span, ̄D h is a sound directivity function and r is the distance to the
observer. The additional terms Fi(St) and Gi(Re) are functions of the ‘Strouhal number’ St ¼ fδ/U and the ‘Reynolds numbe’r Re.
The nature of dependency is different for each noise mechanism, but it is impressive that all the formulae look so much alike.
For turbulent inflow, the prediction equation is normally based on the work of Amiet (Sørensen, 2009). This model takes the
following form:
     
DL 2 7=3 Kc
Lp ¼ 10 log r20 c20 l 2 M3 I2 ^k 1 þ ^k
3
þ 58:4 þ 10 log ; ð61Þ
r 1 þ Kc
where l is a turbulence length scale, I is the turbulence intensity, r0is the density, c0the speed of sound, DL is the blade segment
semi-span, ^k is a corrected wave length and Kc is a low frequency correction.
Taking into account all the variable dependencies, the problem of predicting the noise spectrum at a given observer's 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 un-tripped 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.

In the present study the theory behind the empirical correlations is not studied in detail and for details about the nature of each
of the modeled noise mechanisms the reader may consult the original work of Brooks et al. (Brooks et al., 1989) and Amiet
(Amiet, 1975) or the textbook of Wagner et al. (Wagner et al., 1996).
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 (Drela, 1989). 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's position are calculated for each noise mechanism and for each blade element.
Aerodynamic Analysis of Wind Turbines 191

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 semi-empirical 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

Ainslie JF (1985) Development of an eddy viscosity model for wind turbine wakes. Proc.7th BWEA Wind Energy Conference Oxford.
Amiet RK (1975) Acoustic radiation from an airfoil in a turbulent stream. Journal of Sound and Vibration 41: 407–420.
Bangga G, Kim T, Lutz T, Weihing P, and Krämer E (2016) Investigations of the inflow turbulence effect on rotational augmentation by means of CFD. Journal of Physics:
Conference Series 753(2): 022026.
Bangga G, Lutz T, Jost E, and Krämer E (2017) CFD studies on rotational augmentation at the inboard sections of a 10 MW wind turbine rotor. Journal of Renewable and
Sustainable Energy 9. 023304.
Betz A (1920) Das Maximum der theoretisch möglichen Ausnützung des Windes durch Windmotoren. Zeitschrift für das gesamte Turbinenwesen 26: 307–309.
Breton SP, Sumner J, Sørensen JN, Hansen KS, Sarmast S, and Ivanell S (2017) A survey of modelling methods for high-fidelity wind farm simulations using large eddy
simulation. Philosophical Transactions of the Royal Society A 375: 20160097.
Brooks TF, Pope DS, and Marcolini MA (1989) Airfoil self-noise and prediction. NASA Reference Publication 1218, USA: National Aeronautics and Space Administration.
Calderer A, Guo X, Shen L, and Sotiropoulos F (2014) Coupled fluid-structure interaction simulation of floating offshore wind turbines and waves: A large eddy simulation
approach. IOP Journal of Physics: Conference Series 524(1):
Chaviaropoulos PK, Nikolaou IG, Aggelis K, Sørensen NN, Montgomerie B, von Geyr H, Hansen MOL, Kang S, Voutsinas S, and Dyrmose SZ (2001) Viscous and aeroelastic
effects on wind turbine blades: The VISCEL project. Proc. 2001 European Wind Energy Conference and Exhibition, Munchen.
Chehouri A, Younes R, Ilinca A, and Perron J (2015) Review of performance optimization techniques applied to wind turbines. Applied Energy 142: 361–388.
Churchfield MJ, Schreck SJ, Martinez LA, Meneveau C, and Spalart PR (2017) An advanced actuator line method for wind energy applications and beyond. Presented at the
American Institute of Aeronautics and Astronautics Sci Tech 2017, Grapevine, Texas, January 9–13.
Cottet G-H and Koumoutsakos PD (2000) Vortex methods: Theory and Practice. Cambridge University Press.
Crespo A and Hernández J (1996) Turbulence characteristics in wind-turbine wakes. Journal of Wind Engineering and Industrial Aerodynamics 61(1): 71–85.
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.
Dag KO (2017) Combined pseudo-spectral/actuator line model for wind turbine applications. PhD thesis Technical University of Denmark.
Dag KO and Sørensen JN (2020) A new tip correction for actuator line computations. Wind Energy 23(2): 148–160.
Draper M, Guggeri A, and Usera G (2016) Validation of the actuator line model with coarse resolution in atmospheric sheared and turbulent inflow. IOP Journal of Physics:
Conference Series 753(8):
Drela M (1989) XFOIL: An analysis and design system for low Reynolds number airfoils. Conf. on Low Reynolds Number AerodynamicsUniversity Notre Dame.
Duque EPN, van Dam CP, and Hughes S (1999) Navier-stokes simulations of the NREL combined experiment phase II rotor. AIAA Paper 99-0037.
Duque EP, Burklund MD, and Johnson W (2003) Navier-stokes and comprehensive analysis performance predictions of the NREL phase VI experiment. Journal
of Solar Energy Engineering 125(4): 457–467.
Forsting ARM, Pirrung GR, and Ramos-Garcia N (2020) A fast vortex-based smearing correction for the actuator line. Wind Energy Science 5(1): 349–353.
Frandsen S, Barthelmie R, Pryor S, Rathmann O, Larsen S, Højstrup J, and Thøgersen M (2006) Analytical modelling of wind speed deficit in large offshore wind farms. Wind
Energy 9: 39–53.
Froude W (1878) On the elementary relation between pitch, slip and propulsive efficiency. Transactions of the Royal Institution of Naval Architects 19: 47.
Froude RE (1889) On the part played in propulsion by difference of fluid pressure. Trans. Roy. Inst. Naval Arch. 30: 390–405.
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).
Fuglsang P and Madsen HA (1999) Optimization method for wind turbine rotors. Journal of Wind Engineering and Industrial Aerodynamics 80: 191–206.
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.
Glauert H (1935) In: Durand W.F. (ed.), Airplane Propellers. Division L in Aerodynamic Theory, vol. IV, pp. 169–360. Berlin: Springer.
Grinderslev C, Horcas SG, and Sørensen NN (2021) Fluid-structure interaction simulations of a wind turbine rotor in complex flows, validated through field experiments. Wind
Energy. doi:10.1002/we.2639.
Guma G, Bangga G, Lutz T, and Krämer E (2021) Aeroelastic analysis of wind turbines under turbulent inflow conditions. Wind Energy Science 6(1): 93–110.
Guntur S and Sørensen NN (2015) A study on rotational augmentation using CFD analysis of flow in the inboard region of the MEXICO rotor blades. Wind Energy 18(4):
745–756.
Hansen MOL (2008) Aerodynamics of Wind Turbines. London: Earthscan.
Hansen MOL, Sørensen JN, Voutsinas S, Sørensen N, and Madsen HA (2006) State of the art in wind turbine aerodynamics and aeroelasticity. Progress in Aerospace Science
42: 285–330.
Ivanell SA (2009) Numerical computations of wind turbine wakes. PhD dissertation Stokholm: KTH, Royal Institute of Technology.
Ivanell S, Sørensen JN, Mikkelsen R, and Henningson D (2008) Analysis of numerically generated wake structures. Wind Energy 12(1): 63–80.
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.
Jensen N (1983) A note on wind generator interaction. Technical Report Risø-M-2411, Risø.
Jha PK and Schmitz S (2018) Actuator curve embedding - an advanced actuator line model. Journal of Fluid Mechanics 834(R2):
Jha PK, Churchfield MJ, Moriarty PJ, and Schmitz S (2014) Guidelines for volume force distributions within actuator line modeling of wind turbines on large-Eddy simulation-
type grids. Journal of Solar Energy Engineering 136(3):
Johansen J and Sørensen NN (2004) Aerofoil characteristics from 3D CFD rotor computations. Wind Energy 7(4): 283–294.
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.
Joukowsky NE (1912) Vortex theory of a rowing screw. Trudy Otdeleniya Fizicheskikh Nauk Obshchestva Lubitelei Estestvoznaniya 16(1): 1–31.
Katz J and Plotkin A (1991) Low-Speed Aerodynamics. New York: McGraw-Hill.
Khlaifat N, Altaee A, Zhou J, and Huang Y (2020) A review of the key sensitive parameters on the aerodynamic performance of a horizontal wind turbine using computational
fluid dynamics modelling. AIMS Energy 8(3): 493–524.
Lanchester FW (1915) A contribution to the theory of propulsion and the screw propeller. Transactions of the Institution of Naval Architects 57: 98.
Larsen GC (1988) A simple wake calculation procedure. Technical Report Risø-M-2760, Risø.
Leclerc C and Masson C (2004) Towards blade-tip vortex simulation with an actuator-lifting surface model. AIAA-2004-0667.
192 Aerodynamic Analysis of Wind Turbines

Leishman JG (2002) Challenges in modelling the unsteady aerodynamics of wind turbines. Wind Energy 5: 85–132.
Leishman JG and Beddoes TS (1989) A semi-empirical model for dynamic stall. Journal of the American Helicopter Society 34(3): 3–17.
Lesieur M, Metais O, and Comte P (2005) Large-Eddy Simulations of Turbulence. Cambridge University Press.
Li Y, Castro AM, Sinokrot T, Prescott W, and Carrica PM (2015) Coupled multi-body dynamics 8and CFD for wind turbine simulation including explicit wind turbulence.
Renewable Energy 76: 338–361.
Lutz T, Herrig A, Würz W, Kamruzzaman M, and Krãmer E (2007) Design and wind-tunnel verification of low-noise airfoils for wind turbines. AIAA Journal 45(4): 779–785.
Margoulis W (1922) Propeller theory of Professor Joukowski and his pupils. Memor: NACA Tech, No. 79.
Marten D, Wendler J, Pechlivanoglou G, Nayeri CN, and Paschereit CO (2013) QBlade: An open source tool for design and simulation of horizontal and vertical Axis wind
turbines. International Journal of Emerging Technology and Advanced Engineering 3(3): 264–269.
Marten D, Lennie M, Pechlivanoglou G, Nayeri CN, and Paschereit CO (2016) Implementation, optimization, and validation of a nonlinear lifting line-free Vortex wake module
within the wind turbine simulation code QBLADE. Journal of Engineering for Gas Turbines and Power 138(7):
Martinez-Tossas LA and Meneveau C (2019) Filtered lifting line theory and application to the actuator line model. Journal of Fluid Mechanics 863: 269–292.
Martinez-Tossas LA, Churchfield MJ, and Meneveau C (2017) Optimal smoothing length scale for actuator line models of wind turbine blades based on Gaussian body force
distribution. Wind Energy 20(6): 1083–1096.
Mehta D, van Zuijlen A, Koren B, Holierhoek J, and Bijl H (2014) Large eddy simulation of wind farm aerodynamics: A review. Journal of Wind Engineering and Industrial
Aerodynamics 133: 1–17.
Menter FR (1993) Zonal two-equation k  omodels for aerodynamic flows. AIAA Paper 93-2906.
Mikkelsen R (2003) Actuator disc methods applied to wind turbines. PhD dissertation Lyngby, Denmark: DTU Mechanical Engineering.
Moriarty P, Guidati G, and Migliore P (2005) Prediction of Turbulent Inflow and Trailing-Edge Noise for Wind Turbines. AIAA Paper 2005-2881, Proc. of the 11th AIAA/CEAS
Aeroacoustics Conf., Monterey.
Nilsson K, Ivanell S, Hansen KS, Mikkelsen R, Sørensen JN, Breton S-P, and Henningson D (2015) Large-eddy simulations of the Lillgrund wind farm. Wind Energy 18:
449–467.
Okulov VL and Sørensen JN (2008) Refined Betz limit for rotors with a finite number of blades. Wind Energy 11: 415–426.
Okulov VL and Sørensen JN (2010) Maximum efficiency of wind turbine rotors using Joukowsky and Betz approaches. Journal of Fluid Mechanics 649: 497–508.
Øye S (1991) Dynamic stall, simulated as a time lag of separation. Proc. 4th IEA Symposium on the Aerodynamics of Wind Turbines. ETSU-N-118, Harwell, U.K.
Pape L and Lecanu J (2004) 3D Navier–stokes computations of a stall-regulated wind turbine. Wind Energy 7(4): 309–324.
Pierella F, Krogstad P-Å., and Sætran L (2014) Blind test 2 calculations for two in-line model wind turbines where the downstream turbine operates at various rotational
speeds. Renewable Energy 70: 62–77.
Porté-Agel F, Bastankhah M, and Shamsoddin S (2020) Wind-turbine and wind-farm flows: A review. Boundary-Layer Meteorology 174: 1–59.
Ramos García N, Hejlesen MM, Sørensen JN, and Walther JH (2017) Hybrid vortex simulations of wind turbines using a three-dimensional viscous-inviscid panel method.
Wind Energy 20(11): 1871–1889.
Ramos-Garcia N, Shen WZ, and Sørensen JN (2014a) A strong viscous-inviscid interaction model for rotating airfoils. Wind Energy 17: 1957–1984.
Ramos-Garcia N, Sørensen JN, and Shen WZ (2014b) Validation of a three-dimensional viscous-inviscid interactive solver for wind turbine rotors. Renewable Energy 70:
78–92.
Ramos-García N, Sørensen JN, and Shen WZ (2016) Three-dimensional viscous-inviscid coupling method for wind turbine computations. Wind Energy 19(1): 67–93.
Rankine WJM (1865) On the mechanical principles of the action of propellers. Transactions of the Institution of Naval Architects 6: 13.
Sanderse B, Pijl S, and Koren B (2011) Review of computational fluid dynamics for wind turbine wake aerodynamics. Wind Energy 14(7): 799–819.
Santo G, Peeters M, Van Paepegem W, and Degroote J (2020) Effect of rotor-tower interaction, tilt angle, and yaw misalignment on the aeroelasticity of a large horizontal axis
wind turbine with composite blades. Wind Energy 23(7): 1578–1595.
Sarlak H, Meneveau C, and Sørensen JN (2015) Role of subgrid-scale modeling in large eddy simulation of wind turbine wake interactions. Renewable Energy 77: 386–399.
Schepers JG and Snel H (1995) Dynamic inflow: Yawed conditions and partial span pitch control. In: ECN-C-95-056. Petten: The Netherlands.
Schepers JG, Curvers APWM, Oerlemans S, Braun K, Lutz T, Herrig A, Wuerz W, and Mendez Lopez B (2005) SIROCCO: SIlent ROtors by aCoustic Optimalisation. 1st Int.
Meet. on Wind Turbine Noise: Perspectives for Control.
Schreck S (2002) The NREL full-scale wind tunnel experiment introduction to the special issue. Wind Energy 5(2–3): 77–84.
Schubel PJ and Crossley RJ (2012) Wind Turbine Blade Design. Energies 5(9): 3425–3449.
Schulz C, Letzgus P, Lutz T, and Krämer E (2017) CFD study on the impact of yawed inflow on loads, power and near wake of a generic wind turbine. Wind Energy 20:
253–268.
Shen WZ, Mikkelsen R, Sørensen JN, and Bak C (2005) Tip loss corrections for wind turbine computations. Wind Energy 8(4): 457–475.
Shives M and Crawford C (2013) Mesh and load distribution requirements for actuator line CFD simulations. Wind Energy 16(8): 1183–1196.
Snel H and Schepers JG (1995) Joint Investigation of Dynamic Inflow Effects and Implementation of an Engineering Method. In: ECN-C-94-107. Petten: The Netherlands.
Snel H and van Holten T (1995) Review of recent aerodynamic research on wind turbines with relevance to rotorcraft. AGARD Report CP-552, Chap. 7, pp. 1–11.
Viscwind: Viscous effects on wind turbine blades. In: Sørensen JN (ed.), Report ET-AFM-9902. Lyngby, Denmark: Department of Energy Engineering, DTU.
Sørensen NN (2009) CFD modelling of laminar-turbulent transition for airfoils and rotors using the g  R~ ey model. Wind Energy 12(8): 715–733.
Sørensen JN (2011) Aerodynamic aspects of wind energy conversion. Annual Review of Fluid Mechanics 43: 427–448.
Sørensen NN and Hansen MOL (1998) Rotor performance predictions using a Navier-Stokes method. AIAA Paper 98-0025.
Sørensen JN and Kock CW (1995) A model for unsteady rotor aerodynamics. Journal of Wind Engineering and Industrial Aerodynamics 58: 259–275.
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.
Sørensen JN, Dag KO, and Ramos-Garcia N (2016) A refined tip correction based on decambering. Wind Energy 19(5): 787–802.
Sørensen JN and Shen WZ (2002) Numerical modelling of wind turbine wakes. Journal of Fluids Engineering 124(2): 393–399.
Sørensen NN, Michelsen JA, and Schreck S (2002) Navier-Stokes predictions of the NREL phase VI rotor in the NASA-AMES 80 ft x 120 ft wind tunnel. Wind Energy 5:
151–169.
Sørensen JN, Hansen MOL and Hansen KS (eds.), 2007. The Science of Making Torque from Wind. Journal of Physics Conference Series 75.
Sørensen NN, Bechmann A, Réthoré PE, and Zahle F (2014) Near wake Reynolds-averaged Navier–stokes predictions of the wake behind the MEXICO rotor in axial and yawed
flow conditions. Wind Energy 17(1): 75–86.
Spalart P and Allmaras S (1994) A one-equation turbulence model for aerodynamic flows. La Recherche Aérospatiule 1(1): 5–21.
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.
Spera DA (1994) Wind Turbine Technology. New York: ASME Press.
Stevens RJAM and Meneveau C (2017) Flow structure and turbulence in wind farms. Annual Review of Fluid Mechanics 49(1): 311–339.
Stevens RJAM, Martínez-Tossas LA, and Meneveau C (2018) Comparison of wind farm large eddy simulations using actuator disk and actuator line models with wind tunnel
experiments. Renewable Energy 116: 470–478.
Theodorsen T (1948) Theory of Propellers. New York: McGraw-Hill Book Company.
Tran TT and Kim DH (2018) A CFD study of coupled aerodynamic-hydrodynamic loads on a semisubmersible floating offshore wind turbine. Wind Energy 21(1): 70–85.
Aerodynamic Analysis of Wind Turbines 193

Troldborg N (2008) Actuator line modelling of wind turbine wakes. PhD dissertation Lyngby, Denmark: DTU Mechanical Engineering.
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.
Troldborg N, Larsen GL, Madsen HA, Hansen KS, Sørensen JN, and Mikkelsen R (2010) Numerical simulations of wake interaction between two wind turbines at various inflow
conditions. Wind Energy. doi:10.1002/we.433.
van Kuik GAM (2007) The Lanchester-Betz-Joukowsky limit. Wind Energy 10: 289–291.
Vermeer LJ, Sørensen JN, and Crespo A (2003) Wind turbine wake aerodynamics. Progress in Aerospace Science 39: 467–510.
Viterna LA and Corrigan RD (1981) Fixed-pitch rotor performance of large HAWT’s. DOE/NASA Workshop on Large HAWTs. Cleveland, Ohio.
Voutsinas SG (2006) Vortex methods in aeronautics: How to make things work. International Journal of Fluid Dynamics 20(1): 3–18.
Wagner S, Bareiss R, and Guidati G (1996) Wind Turbine Noise. EUR 16823 Springer.
Wilson RE and Lissaman (1974) P.B.S. Applied Aerodynamics of Wind Power Machines. Oregon State University.
Wind Turbine Aerodynamics: A State-of-the-Art. In: Brouckaert J-F (ed.), VKI Lecture Series 2007–05. Belgium: von Karman Institute for Fluid Dynamics.
Wu YT and Porté-Agel F (2011) Large-eddy simulation of wind-turbine wakes: Evaluation of turbine parametrisations. Boundary-Layer Meteorology 138(3): 345–366.
Wu YT and Porté-Agel F (2013) Simulation of turbulent flow inside and above wind farms: Model validation and layout effects. Boundary-Layer Meteorology 146(2): 181–205.
Xie S and Archer C (2015) Self-similarity and turbulence characteristics of wind turbine wakes via large-eddy simulation. Wind Energy 18: 1815–1838.
Xudong W, Shen WZ, Zhu WJ, Sørensen JN, and Jin C (2009) Shape optimization of wind turbine blades. Wind Energy 12(8): 781–803.
Yang X, Sotiropoulos F, Conzemius RJ, Wachtler JN, and Strong MB (2014) Large-eddy simulation of turbulent flow past wind turbines/farms: The virtual wind simulator
(VWiS). Wind Energy 18: 2025–2045.
Yu DO and Kwon OJ (2014) Time-accurate aeroelastic simulations of a wind turbine in yaw and shear using a coupled CFD-CSD method. Journal of Physics: Conference
Series 524(1): 12046.
Zahle F, Sørensen NN, and Johansen J (2009) Wind turbine rotor-tower interaction using an incompressible overset grid method. Wind Energy 12(6): 594–619.
Zhu WJ, Heilskov N, Shen WZ, and Sørensen JN (2005a) Modeling of aerodynamically generated noise from wind turbines. Journal of Solar Energy Engineering 127: 517–528.
Zhu WJ, Sørensen JN, and Shen WZ (2005b) An aerodynamic noise propagation model for wind turbines. Wind Engineering 29(2): 129–143.

Further reading

Burton T, Sharpe D, Jenkins N, and Bossanyi E (2001) Wind Energy Handbook. New York: Wiley.
De Vries O (1979) Fluid Dynamic Aspects of Wind Energy Conversion. AGARD Report AG-243.
Hansen AC and Butterfield CP (1993) Aerodynamics of horizontal-axis wind turbines. Annual Review of Fluid Mechanics 25: 115–149.
Hansen MOL (2008) Aerodynamics of Wind Turbine. London: Earthscan.
Hansen MOL, Sørensen JN, Voutsinas S, Sørensen N, and Madsen HA (2006) State of the art in wind turbine aerodynamics and aeroelasticity. Progress in Aerospace Science
42: 285–330.
Hau E and von Renouard H (2006) Wind Turbines: Fundamentals, Application, Economics. Berlin: Springer Verlag.
Leishman JG (2012) Challenges in modeling the unsteady aerodynamics of wind turbines. Wind Energy 5: 86–132.
Manwell F, McGowan JG, and Rogers AL (2010) Wind Energy Explained: Theory, Design and Application. John Wiley & Sons.
Schaffarczyk AP (2014) Understanding Wind Power Technology: Theory, Deployment and Optimisation. John Wiley & Sons.
Schmitz S (2020) Aerodynamics of Wind Turbines: A Physical Basis for Analysis and Design. John Wiley & Sons.
Snel H (1998) Review of the present status of rotor aerodynamics. Wind Energy 1: 46–69.
Sørensen JN (2011) Aerodynamic aspects of wind energy conversion. Annual Review of Fluid Mechanics 43: 427–448.
Sørensen J.N. and Sørensen J.D. (eds.), 2011. Wind energy systems: Optimising design and construction for safe and reliable operation. Woodhead Publishing Series in
Energy, vol. 10. Cambridge.
Sørensen JN (2016) General momentum theory for horizontal axis wind turbines. In: Research Topics in Wind Energy, vol. 4. Springer.
Spera DA (1994) Wind Turbine Technology. New York: ASME Press.
Vermeer LJ, Sørensen JN, and Crespo A (2003) Wind turbine wake aerodynamics. Progress in Aerospace Science 39: 467–510.

You might also like