You are on page 1of 13

Composite Structures 159 (2017) 144–156

Contents lists available at ScienceDirect

Composite Structures
journal homepage: www.elsevier.com/locate/compstruct

Proposal for a coupled aerodynamic–structural wind turbine blade


optimization
Andrea Dal Monte ⇑, Stefano De Betta, Marco Raciti Castelli, Ernesto Benini
Department of Industrial Engineering, University of Padua, Via Venezia 1, I-35131 Padua, Italy

a r t i c l e i n f o a b s t r a c t

Article history: An advanced design algorithm for the AOC 15/50 wind turbine is hereby presented: both aerodynamic
Received 31 May 2016 and structural parameters are considered as the design variables for a coupled BEM–FEM optimization.
Revised 15 September 2016 In order to obtain an improvement in rotor blade design, a modified version of the S.O.C.R.A.TE. algo-
Accepted 15 September 2016
rithm combines a BEM evaluation of its aerodynamic performance, a FEM structural analysis and a
Available online 17 September 2016
genetic algorithm. Both the aerodynamic power (resulting from the BEM analysis) and the tip displace-
ment (produced by the aerodynamic forces and the inertial loads) are considered as the fitness functions
Keywords:
for the optimization problem.
HAWT
FEM analysis
An improved distribution of both chord values and twist angles is determined, as well as an optimal
BEM theory layout of the blade composite skin. A slight increase in the aerodynamic power generation is obtained,
Evolutionary algorithm as well as a marked improvement in overall blade deformation characteristics.
S.O.C.R.A.TE. algorithm Ó 2016 Elsevier Ltd. All rights reserved.

1. Introduction and background aerodynamic performance of a 2.2 m diameter three-bladed HAWT


in order to assess the applicability of the BEM Theory for the mod-
The horizontal-axis wind turbine (HAWT) represents the most elling of small scale rotors. ElQuatary and Elhadidi [4] compared
common architecture among existing wind energy conversion sys- BEM and Computational Fluid Dynamics (CFD) simulations for
tems, with thousands of MWs of new capacity worldwide installed two HAWTs characterized by different blade thickness, registering
each year. Its design process, largely accepted by manufacturers as a marked agreement especially for the thicker blade configuration.
well as by academic institutions, is generally separated in two con- Kong et al. [5] proposed a structural design of a medium scale com-
secutive stages [1]: posite HAWT blade made of E-glass/epoxy. Several design load
cases (such as aerodynamic forces, those due to ice accumulation,
 the external geometry of the blade (in terms of both chord and hygro-thermal and mechanical loads) were considered and the
twist angle distribution along the blade span, rotor size and most dominant design parameters were included in a FEM analy-
other factors, often empirical, related to the cost of energy) is sis, also estimating the fatigue life of the blade.
first determined using a Blade Element Momentum (BEM) Among numerical optimization methods, particular relevance is
based algorithm nowadays assumed by evolutionary algorithms, whose solutions
 a proper layout of both blade skin and reinforcements is deter- are generated on the basis of techniques inspired by natural evolu-
mined by means of a structural analysis based on the finite ele- tion. As observed by Mendez and Greiner [6], genetic algorithms
ment method (FEM), considering both the aerodynamic and are global optimizers that have a wide trade-off between explo-
inertial loads acting on the blade. ration and exploitation of the space problem: among their advan-
tages, a global search capability is to be recognized, due to the
Both stages have been widely investigated by several authors. management of a population of candidate solutions instead of only
Liu and Janajreh [2] proposed an improved BEM model for the one. Moreover, their only requirement is the knowledge of the fit-
analysis of HAWT performance, considering both the tip loss effect ness function, without any other consideration such as its deriv-
and the rotational one, with the aim of extending its application to ability or continuity. A great number of engineering problems
the turbulent wake regime. Refan and Hangan [3] investigated the can be dealt with genetic algorithms [7]: Benini and Toffolo [8]
performed a multi-objective optimization for the design of stall-
regulated HAWTs, coupling the BEM Theory and a multi-
⇑ Corresponding author.
objective evolutionary algorithm, with the scope of achieving the
E-mail address: andrea.dalmonte@gmx.com (A. Dal Monte).

http://dx.doi.org/10.1016/j.compstruct.2016.09.042
0263-8223/Ó 2016 Elsevier Ltd. All rights reserved.
A. Dal Monte et al. / Composite Structures 159 (2017) 144–156 145

Nomenclature

A ½m2  blade surface m [m] blade mass


A0 ½m2  original AOC 15/50 blade surface m0 [m] mass of the original AOC 15/50 blade
c ½m section chord length M [Nm] bending moment acting on each blade cross section
d ½mm total displacement at the tip p [–] generation number
d0 ½mm total displacement at the tip of the original AOC 15/50 pmax [–] maximum generation number
blade P ½kW aerodynamic power generated from the rotor
EI ½MN  m2  blade flexural rigidity P0 ½kW aerodynamic power generated from the original AOC
EL ½MPa longitudinal elastic modulus 15/50 rotor
ET ½MPa transversal elastic modulus r [m] sectional distance from the axis of rotation
EZ ½MPa out-of-plane elastic modulus R [m] blade radius
F ½N aerodynamic force acting on the blade section t [mm] thickness of the layer
f A [–] ratio between blade surface area and that of the original V [m/s] unperturbed wind velocity
AOC 15/50 blade a ½  section angle of attack
f P [–] objective function of power h ½  section twist angle
dH
f d [–] objective function of deformation dz
[–] rate of rotation of the blade section
f m [–] ratio between blade mass and that of the original AOC mLT [–] in plane Poisson’s ratio
15/50 blade mTZ [–] out-of-plane Poisson’s ratio
GLT ½MPa in-plane shear modulus mLZ [–] out-of-plane Poisson’s ratio
GTZ ½MPa out-of-plane shear modulus q ½kg=m3 
GLZ ½MPa out-of-plane shear modulus material density
iA [–] penalty function for the area qA ½kg=m3  blade surface density
im [–] penalty function for the mass
iP [–] penalty function for the power

best trade-off between annual energy production per square metre problem but using a combined cost (AEP divided by total weight),
and cost of energy. Cai et al. [9] developed a structural optimiza- a levelized cost of energy or the Cost Of Energy only.
tion of an HAWT blade using a particle swarm optimization algo- Even though aerodynamic and structural optimizations of
rithm based on FEM calculations, proving a great potential HAWT blades have been widely proposed by several authors, in
improvement on overall structural blade performance. Dal Monte reviewing the literature, the potential of an evolutionary algorithm
et al. [10] improved the structural response of the AOC 15/50 San- based on the coupling of an aerodynamic model (based upon the
dia blade using the S.O.C.R.A.TE. (Structural Optimization for Com- BEM Theory) and a structural one (based on a FEM analysis) have
posite Rotor Air TurbinE) algorithm: both the choice of the been not often investigated; Zhu et al. [16] proposes an aerody-
employed materials and their placement in the layout of the blade namic and structural integrated optimization for the HAWT Blades
skin were considered as design variables for the optimization, design, Wang et al. [17] developed an aerodynamic and structural
obtaining a marked reduction in the mass of the blade and a corre- integrated design optimization method for a composite wind tur-
sponding increment of its flapwise rigidity. An optimization proce- bine blade based on multidisciplinary design optimization (MDO).
dure for a HAWT blade based upon an ultimate limit state analysis Gradient-based optimizers have also proved their capabilities in
was proposed by Hu et al. [11]: in order to minimize the blade cost aerospace optimization. They have played and continue to play a
and its total mass, two different composite materials, such as glass key role during the aero-structural design of the aircraft. Ghom-
fibre reinforced plastic (GFRP) and carbon fibre reinforced plastic mem et al. [18] implemented a shape optimization of flapping
(CFRP) were considered, being the design variables of the blade wings in forward flight, combining a gradient-based optimizer
skin the input parameters for a combined FEM and evolutionary (GCMMA) with the unsteady vortex lattice method (UVLM). Gille-
algorithm analyses. bart et al. [19] presented a two-dimensional low-fidelity aero-
Several tools for the multi-disciplinary wind turbine optimiza- elastic analysis of an airfoil and a gradient based optimization
tion have been proposed in the open literature in the last years; (GCMMA) consisting of a coupled potential flow model and curved
Pourrajabian et al. [12] proposed a procedure for the aero- Timoshenko beam model combined with a boundary layer model.
structural design of a small wind turbine blade based on a BEM A great advantage of the gradient-based optimizer is to handle a
code and on a simple structural model. Bottaso et al. [13] described large number of design variables and coinstrants; furthermore they
a procedure for the multidisciplinary optimization of wind tur- result faster and less computational expensive compared to genetic
bines with a parametric high fidelity aero-servo-elastic model, algorithms, however a potential weakness is the relative intolerant
considering the Annual Energy Production and the Weight of the of difficulties such as noisy objective function spaces and topology
blade as cost functions. Ashuri et al. [14] also developed a multidis- optimization; additionally they find a local rather then a global
ciplinary optimization for the design of offshore wind turbines; the minimum [20]. The characteristics of the analysed problem poten-
considered objective functions is represented by the levelized cost tially involve several local minimum, furthermore the evaluation of
of energy and it included design constraints as stresses, deflections both aerodynamic and structural function is not expensive in
modal frequencies and fatigue limits. Grujicic et al. [15] proposed a terms of time. For such reason the genetic algorithm formulation
multidisciplinary design optimization procedure for the develop- has been chosen as optimization method over the gradient based
ment of the cost effective composite layout of an HAWT using formulation. The hereby proposed genetic algorithm considers at
the Cost of Energy (COE) as single fitness function. In the cited tools the same time both BEM and FEM genes, in order to determine
the multi-objective design is not formulated as a Pareto optimal an aerodynamic fitness function and a structural one. The purpose
of the present optimization is therefore to increase both the power
146 A. Dal Monte et al. / Composite Structures 159 (2017) 144–156

production and the flapwise rigidity of the blade, using an iterative Table 2
BEM–FEM analysis. Aerodynamic profiles characterizing the external geometry of the AOC 15/50 blade.

Distance from the rotational axis [m] Blade profile c [m] h [°]
2. The case study 0.279 S814 0.4570 8.10
0.775 S814 0.5303 8.10
The AOC 15/50 wind turbine was selected as case study for the 0.88 S814 0.5462 7.60
1.092 S814 0.5745 6.72
present optimization as it represents the validate test case for the
1.702 S814 0.6568 6.24
S.O.C.R.A.TE. software. The turbine is one of the few examples with 2.311 S814 0.7173 5.94
both openly available structural data (with the complete descrip- 3.124 S814 0.7400 5.53
tion of the composite layout of the blade) and experimental results 3.937 S814 0.6920 4.86
4.750 S812 0.6258 4.10
(experimental power curve resulting from a test campaign). Using
5.563 S812 0.5572 3.28
this example of wind turbine is possible to validate both the mod- 6.603 S813 0.4780 2.37
els (structural and aerodynamic) on the same geometry. The main 7.059 S813 0.4410 1.95
characteristics of the AOC 15/50 are reported in the NREL Test 7.490 S813 0.4060 1.54
Report [21] and summarized in Table 1.
The AOC 15/50 HAWT blade was initially designed by the San-
dia National Laboratories [22]: starting at 0.279 m from the rota- others, and according to the classical theory, the force exerted from
tional axis and extending up to 7.490 m, the blade was based on the blades on the flow is considered constant in each annular ele-
the S821, S819 and S820 airfoils (from root to tip). An improved ment. This corresponds to an assumption of a rotor characterized
version of the AOC 15/50 blade [23] [24] is based on the S814, by an infinite blade number [25], which, of course, has a vortex
S812 and S813 airfoils (from root to tip, placed at 7.2 m from the system in the wake that is different from that of an actual rotor
rotational axis): such configuration is adopted in the hereby pro- with a finite number of blades. In order to simulate a rotor with
posed analysis. a finite blade number, the classical Prandtl’s tip loss correction
Table 2 summarizes the main geometrical features of the blade [27] is hereby considered. A detailed description of the process
model, while the layup schedule of the adopted composite materi- with an extensive explanation of its fundamental equations can
als is reported in Table 3: the layer number increases from the tip be found in [25].
to the root. As can be observed from Fig. 1, several reinforcements Among the advantages of such approach with respect to more
are added to the blade skin. Furthermore, in order to increase the advanced tools (like Computational Fluid Dynamics), its high com-
overall blade rigidity, both a Spar Flange and a Spar Web are putational speed and ease of implementation can be recalled [28–
adopted in the central sections (from the spanwise coordinate 30]. However, unlike more advanced calculation methods, the
1.092 m to 7.061 m), as can be seen from Fig. 2. It can also be accuracy of the results is not always ensured and is heavily influ-
observed that the blade is subdivided in 9 main zones in the span- enced by the precision of the airfoil polars characterizing each
wise direction and in 5 chordwise areas. blade station. Moreover, experimental polars are often not avail-
able, requiring the adoption of numerical codes based on panel
3. Description of the BEM code methods, such as XFoil [31], for the prediction of airfoil character-
istics. In this work the RFoil code [32], an improved version of
The proposed BEM code has been developed and validated by XFoil, is used due to its better correlation with experimental
the researchers of the University of Padova (as the S.O.C.R.A.TE. results, especially in the stall region. The AOC 15/50 wind turbine
software) and it is based on the well-known theory presented by is stall regulated and, even using the RFoil code, the stall condition
Martin O.L. Hansen [25]. The intent is to further develop the inter- could not be well captured. However this don’t represents an issue
nal codes in order to archive a complete multidisciplinary analysis for the proposed optimization; the validation of the code shows a
of a horizontal wind turbine. The BEM Theory, whose equations are good prediction even in the stall zone of the power curve. Further-
far too known to be reported here again, is commonly used by more, the aerodynamic objective function refers to the power pro-
wind turbine designers for the prediction of rotor aerodynamic duction in a single point (10 m/s), quite far from the stall
performance. It combines two independent approaches: the conditions as shown in Fig. 3.
Momentum Theory and the Blade Element Theory. The former con- Another weakness of BEM codes is that aerodynamic data are
cerns the computation of both thrust and torque by applying the often available only for a limited range of angles of attack, requir-
conservation of the linear momentum and of the angular one to ing the extension of both lift and drag coefficients up to 90 . This is
a control volume, the latter refers to the analysis of aerodynamic particularly critical for stall controlled wind turbines to accurately
forces acting at each blade section, as a function of blade geometry capture the behaviour of the rotor subjected to high wind speeds.
[26]. Following the BEM method, the rotor is subdivided into a The authors observe that a good agreement with measured data
finite number of control volumes, each independent from the can be obtained following the extension method proposed by Lin-
denburg [33]. In addition, in order to obtain accurate rotor sec-
tional lift characteristics, and hence accurate power prediction
Table 1 (in particular in the stalled region), bi-dimensional airfoil data
AOC 15/50 wind turbine characteristics. need to be corrected for the three-dimensional inboard stall delay
Number of blades 3 effects [34]. In fact, due to rotation, the boundary layer is subjected
Rated power 50 kW to Coriolis and centrifugal forces which alter the bi-dimensional
Cut-in wind speed 4.9 m/s airfoil characteristics [35]. Accordingly, the boundary layer is less
Cut-out wind speed 22.3 m/s
thick and more stable compared to the non-rotational state [33],
Rated wind speed 12 m/s
Rotor diameter 15 m
enhancing the performance of the blades, particularly in their
Online rotational speed 65 rmp innermost portion. To take into account these effects, the correc-
Control type Constant speed – fixed pitch tion on the lift and drag coefficients proposed by Lindenburg [33]
Pitch setting 1.54 is hereby considered.
Power regualation Stall regulation
Tower height 24.4 m
A. Dal Monte et al. / Composite Structures 159 (2017) 144–156 147

Table 3
Layup schedule of the analyzed rotor blade.

Component Location [mm] Layup schedule Thickness [mm] Layup n. [–]


Root 279–775 [45/06 /45/06 /+45]S 15.7 1
775–889 [45/05 /45/05 /+45]S 13.5 2
889–1092 [45/04 /45/04 /+45]S 11.2 3
Spar cap 1092–2311 [45/04 /45/04 /+45]S 11.2 3
2311–3937 [45/03 /45/03 /+45]S 8.9 4
3397–5563 [45/02 /45/02 /+45]S 6.6 5
5563–7493 [45/0/45/0/+45]S 4.3 6
Leading edge 1092–2311 [45/02 /45]S 3.9 7
2311–7493 [45/0/45]S 2.8 8
Trailing edge 1092–6604 [45/0/balsa/0/45] 11.5 9
6004–7493 [45/0]S 2.0 10
Spar flange 1092–7493 [45/02 /45]S 3.9 11
Spar web 1092–7061 [45/02 /45]S 3.9 12

Fig. 3. Experimental and numerical power production as a function of the wind


speed for the AOC 15/50 wind turbine.

ator efficiency is assumed constant and equal to 89:4% (from:


[24]). Fig. 3 shows a comparison between the power curve at sea
level air density and the numerical simulations. The numerical
power curve at low velocities presents an overestimation of the
Fig. 1. Exploded drawing of the analyzed rotor blade geometry, showing the wind turbine performances. This is probably caused by the inertia
subdivision in 9 areas spanwise and 8 zones chordwise. of the wind turbine in the start-up phase; the cut-in speed is
indeed 4.8 m/s and the inertia of the blades (that are starting to
moving) could influence the experimental measurement at low
velocities. A remarkable agreement can be observed up to a wind
velocity of 18 m/s, where a deep stall condition is experienced by
the blades. However, this portion of the power curve can be
neglected for the scope of the present analysis and the accuracy
of the prediction can be therefore considered acceptable.

4. Description of the structural model

The FEM model adopted for the present optimization is built


Fig. 2. Scheme of the subdivision of the blade layout in the chordwise direction.
using SHELL 181 elements to simulate the composite skin of the
blade; such element can reproduce the behaviour of the layered
structures by specifying the sequence of layers, the thickness and
3.1. BEM code validation the orientation of each single lamina and the adopted material.
As shown in Fig. 4, each element is composed of 4 nodes and has
The adopted BEM code is validated against experimental data of 6 degrees of freedom at each node: translations in the nodal x, y
the AOC 15/50 wind turbine installed at NREL’s National Wind and z directions and rotations about the nodal x, y and z axes. A free
Technology Center (NWTC) in Colorado [21]. All the numerical sim- mesh topology and a quad shape of the elements are adopted to
ulations are conduced considering a constant rotor angular speed discretize each surface of the model. A medium dimension of the
of 65 rpm for a range of wind velocities between 5 m/s and surface elements of 20 mm is imposed; a representation of the
20 m/s. Air density is set to 1.225 kg=m3 and the dynamic viscosity mesh in the root area is shown in Fig. 5.
to 1.78 Pa  s. The polar curves for the S814, S812 and S813 airfoils The layup schedule of the model blade is assumed by the Sandia
are obtained using RFoil for a constant Reynolds number of 106 and report [22] and summarized in Table 3. A Layup number is assigned
extended up to 90 using the Lindenburg method [33]. The gener- to every area of the blade where the layup changes. The graphic
148 A. Dal Monte et al. / Composite Structures 159 (2017) 144–156

A130, while DB120 is used for the 45° ply layups. In order to min-
imize the probability of buckling, the balsa wood is adopted as a
filler in the sandwich layup of the trailing edge (zone 9).
A rigid constraint is applied to the root area of the blade: the
three spatial displacements and the three rotations are fixed for
the nodes belonging to the surface. The mechanical loads for the
structural model are obtained by an interpolation of the aerody-
namic forces computed from the BEM model.
The validation of the structural model is obtained by comparing
the results of a FEM analysis to the same results provided in the
AOC 15/50 Sandia report [22]. The flapwise, edgewise and torsional
rigidities of the blade (treated as a cantilever beam) with a fixed
load applied at its tip are computed and compared with experi-
mental results. The detailed description of the validation procedure
is provided in [10] and is not reported here again for brevity’s sake.

5. Design variables
Fig. 4. Shell 181 configuration.
The design variables of the optimization problem are composed
by both BEM and FEM parameters:

1. BEM: the y-coordinates of 5 points of a besier curve represent-


ing the distribution of the chord length in the spanwise direc-
tion (Fig. 6);
2. BEM: the y-coordinates of 5 points of a besier curve represent-
ing the distribution of the twist angle in the spanwise direction
(Fig. 7);
3. FEM: the thickness of each lamina composing the blade
reinforcements.

The adoption of continuous functions to describe chord and


twist distribution is a common and established practice in the
aero-structural parametrization of the wind turbine blades. Exam-
ples of this parametization can be found in [37–39]. Twist and
chord distribution are two of the factors that mainly influence
the power production and the structural behaviour of the blade.
Furthermore the manufacturing costs result to be not greatly
Fig. 5. A detail of the adopted in mesh in the root area. affected by choice of the adopted continuous functions with the
relative constraints, considering the blade built using the common
prepreg moulding technique.
representation of the different layup distribution along the blade is
shown in Fig. 1. Number 11 represents the spar flange layers, over-
6. Formulation of the optimization problem
lapped to the central ones from 3 to 6 and number 12 represents
the Spar Web.
The multi-objective optimization problem concerns the mini-
The adopted materials are some varieties of Glass Reinforced
mization of two functions. The first fitness function f P considered
Polyester (GRP): all the lamina are composed by E-glass fibres
for the optimization process is the result of the BEM analysis. It
embedded in a polymer matrix. The layup is modelled as orthotro-
is represented by the ratio between the power P 0 of the original
pic in a given layer with two of the principal material axes in the
Sandia blade and the power P generated by the candidate blade
plane of the shell, as can be seen from the material parameters
configuration. A second fitness function f d is the result of the struc-
listed in Table 4. Different materials are adopted for the layers with
tural analysis. The total deformation (deriving from the combina-
different orientations of the fibres. The 0° layups are made by

Table 4
Structural properties of the materials adopted in the AOC 15/50 blade.

A130 (0 ) DB120 (45 ) Balsa Wood


EL ¼ EX [MPa] 31700 26200 187
ET ¼ EY [MPa] 7580 6550 61
EZ ¼ EZ [MPa] 7580 6550 4070
mLT ¼ mXY [–] 0.32 0.39 0.67
mTZ ¼ mYZ [–] 0.32 0.35 0.01
mLZ ¼ mXZ [–] 0.32 0.32 0.02
GLT ¼ GXY [MPa] 3450 4140 20.3
GTZ ¼ GYZ [MPa] 3100 3720 150
GLZ ¼ GXZ [MPa] 3100 3720 220
q [kg/m3] 1714 1714 153
Fig. 6. Representation of the Bezier points and of the corresponding curve for the
t [mm] 0.571 0.203 9.530
chord spanwise distribution of the AOC 15/50 blade.
A. Dal Monte et al. / Composite Structures 159 (2017) 144–156 149

The deformation fitness function f d is related to the mass of the


considered solution. A penalty function im is applied to f d when the
mass of the considered blade exceeds by 20% the mass of the orig-
inal one. In order to avoid an excessive increase in the area of the
blade and hence an unfair increment of the generated power, a
penalty function iA is also applied to the value of the power fitness
function f P when the calculated area A exceeds by 5% the original
one A0 .
Finally, In order to deeply explore the zone characterized by a
power fitness function f P close to 1, a second analysis is performed
using a more restrictive upper bound: the power fitness function is
limited to a value of 1.5 and a penalty function iP is applied when
Fig. 7. Representation of the Bezier points and of the corresponding curve for the
twist spanwise distribution of the AOC 15/50 blade. such value exceeds 1.25.
A summary of the main features of the optimization problem is
reported in Table 5.
tion of the flapwise deformation, the edgewise one and the axial The constraints adopted for the optimization problem have
one) of the candidate blade d is compared to the deformation of been found through a trial and error test campaign on the code.
the original Sandia model d0 . High displacements could change The optmization parameters have been proper set after several
the aerodynamic performances so the displacement function has test: the penalty function have been introduced to avoid unacept-
been considered preferable to limit the displacements at the tip able increases of mass, area and to avoid the g.a. to do evolve solu-
of the blade. Alternative structural functions could be represented tion with low power production. Furthermore an alternative
by stress and strain values on the root area, however these values approach has also already been partially tested; the minimization
resulted limited and admissible for an HAWT composite blade, of the mass of the blade has been considered as objective in a pre-
which geometry does not change much from the original one, dur- vious optimization of the AOC 15/50 [10]. In the presented opti-
ing the optimization process. mization, objective functions of different field of analysis, as
The choice of using two contrasting functions belonging to two power production and tip displacement (which affects power),
different fields of analysis instead of a combined index such as the are considered.
Cost Of Energy (already investigated for an HAWT optimization in
[37]) guarantee the identification of a Pareto Front and the possi- 7. The multi-objective genetic algorithm
bility to identify different solutions which favour the structural
or the aero-dynamical behaviour of the blade. The adoption of a multi-objective genetic algorithm represents
The optimization problem is formulated as follow: a powerful strategy in order to improve both the aerodynamic per-
P0 formances and the structural behaviour of a rotor blade. The orig-
min f P ðxÞ ¼  iA ðxÞ  iP ðxÞ ð1Þ inal version of the S.O.C.R.A.T.E. (Structural Optimization for
PðxÞ
Composite Rotor Air TurbinE) algorithm [10] is here modified
dðxÞ and adapted in order to consider the results of a BEM analysis as
min f d ðxÞ ¼  im ðxÞ ð2Þ the input for the following computations. In the present formula-
d0
tion, both chord and twist distributions are initially computed
subject to: from the y-coordinates of two Bezier series of 5 points; the coordi-
f P ðxÞ < 3 ð3Þ nates of the blade profiles are successively calculated. Such vari-
ables are used as input parameters for the BEM analysis, in order
f d ðxÞ < 1:2 ð4Þ to evaluate the aerodynamic power P produced by the blade.
The S.O.C.R.A.T.E. algorithm uses the input geometry and the
( output parameters of the BEM analysis (i.e. the aerodynamic force
1 f m ðxÞ 6 1:20
im ðxÞ ¼ 2 ð5Þ distribution on the blade surface) to generate an appropriate series
ð1 þ f m ðxÞ  1:20Þ f m ðxÞ > 1:20 of APDL (Ansys Parametric Design Language) commands. The
( Ansys software rebuilds the blade using the coordinates of 13 pro-
1 f A ðxÞ 6 1:05 files and applies the aerodynamic forces resulting from the BEM
iA ðxÞ ¼ ð6Þ
ð1 þ ðA  A0 ÞÞ 2
f A ðxÞ > 1:05 analysis (interpolated on these profiles) to the structural model.
A further input for the FEM analysis is represented by the layout
8
<1 P0
6 1:25
PðxÞ
iP ðxÞ ¼ ð7Þ
: ð1 þ f P  1:25Þ P0
> 1:25
PðxÞ Table 5
Summary of the main optimization settings.
where f m is the ratio between the mass of the considered blade m
Variables (1) y-coordinates of Bezier points for the chord
and that of the original one m0 and f A is the ratio between the area distribution
of the considered blade A and that of the original one A0 : (2) y-coordinates of Bezier points for the twist distribution
(3) Layout of blade skin reinforcements
mðxÞ
fm ¼ ð8Þ Fitness (1) Power fitness function f P
m0
functions
(2) Deformation fitness function f d
AðxÞ
fA ¼ ð9Þ Constraints Upper bound on f d
A0 Upper bound on f P
The upper bounds of the power fitness function f P and the Penality function im on f d
Penality function iA on f P
deformation fitness function f d are respectively set to 3 and 1.2.
Penality function iP on f P
150 A. Dal Monte et al. / Composite Structures 159 (2017) 144–156

of the blade skin; the thicknesses of the reinforcement laminas are A certain number of the best individuals in every population is
the structural design variables of the FEM step of the optimization expected to be preserved through the generations: this option of
process. Elitism (Elitecount option) improves the effectiveness of the
Fig. 8 shows the position of the 14 reinforcements (letters from algorithm.
A to P) inside the original AOC 15/50 blade. In order to compute the A peculiarity of the S.O.C.R.A.TE. algorithm is to force the cre-
new thicknesses of the laminas, the algorithm assigns an entire ation and mutation functions of the Matlab gamultiobj to assume
value (from 0 to 3) to the 14 thickness factors T f : if the value of integer values for some genes of the genetic pool. Indeed, the genes
T f is 0, the lamina is deleted in the new configuration; if T f is 1, referred to the structural features of the blade (genes from 11 to
the lamina and its thickness are maintained; if T f is 2 or 3, the lam- 24) represent an integer factor for the thickness of the laminas in
ina is duplicated or tripled. Each blade configuration is encoded the blade skin.
through 24 parameters and the complete genetic pool of an indi- A schematic representation of the current version of the S.O.C.R.
vidual is represented in Fig. 9: the BEM genes for the chord varia- A.T.E. algorithm is shown in Fig. 10. For further details about the S.
tion are coloured in purple, the BEM genes for the twist variation O.C.R.A.TE. algorithm, see [10].
are coloured in blue and the FEM genes for the thickness variation The whole evolutive process is represented by the following
of the reinforcements are coloured in orange. flowchart:
An initial blade population is entirely evaluated first in the BEM
model and then in Ansys, opened by Dos commands in batch mode.  p = 0;
For every blade configuration, Matlab generates a series of profiles  initialization of the parent’s population;
for the BEM code and an APDL file, allowing Ansys to create the  while p = pmax ;
corresponding geometry. In order to evaluate the mechanical beha-  evaluation of the fitness functions;
viour of the blade, a mass computation and a static analysis are  Pareto Ranking (35% of individuals from the first front);
successively run.  selection (Stochastic uniform);
After evaluating the power and the deformation fitness func-  recombination (Crossover Scattered);
tions, all the individuals of the population are sorted using the Par-  mutation;
eto ranking. According to the default Matlab Pareto Ranking  elitism (Elitecount);
option, a fraction of 0.35% of individuals is kept on the first Pareto  new population;
front, while the solver selects individuals from higher fronts.  p = p + 1;
The best individuals of the population are selected and com-
bined in accordance to the criteria of natural selection imple- 8. Description of the optimization process
mented in the gamultiobj function of Matlab. In order to sort the
160 individuals of a generation, the Pareto Ranking method is The optimization process looks for the ideal distribution of both
used: the individuals with highest fitness functions have an high- the chord lengths and the twist angles along the blade span and,
est probability to be chosen for the creation process. Using the simultaneously, for the optimal internal layer distribution. The ini-
Stochastic Uniform method, the algorithm creates a line adding tial population is the result of the combination of two different set
some segments proportional to the fitness values of the individu- of individuals. A first set is composed by individuals whose genes
als. The algorithm moves along the resulting line and chooses the include limited variation compared to the baseline blade genes.
individuals referred to the segment it stops. The second set is random initialized through a specific function
In order to mix the genetic pools of the chosen individuals, the within the boundless of the parameters range. In this way is possi-
Crossover function (scattered option) exchanges some genes ble to explore the whole range of feasible solutions and not exclu-
between the individuals using a random binary vector. In the gen- sively the most promising individual; in addition, the effect of the
erated vector, the 1 values represent a gene from the first individ- variation from the baseline configuration are immediately
ual and the 0 values represent a gene from the second one. The computed.
child is generated according to the scheme of Table 6: The BEM section of the genetic pool is elaborated by Matlab in
The genetic diversity is furthermore safeguarded by a Mutation order to generate chord and twist distributions from the y-
function. The option allows to explore different zones of the space coordinates of the Bezier points. The outputs of the BEM analysis
of variables by introducing small random variations in a certain are represented by the generated power (from which the first fit-
number of genes of some individuals. ness function is computed) and by the values of the aerodynamic
forces acting on the blade. The geometry, the constrains and the
loads are imported in Ansys and both a mass analysis and a struc-
tural one are run. In order to compare the performance of different
blade configurations, the total deformation is estimated on a refer-
ence point located at the tip of the blade (at 30% of the chord
length).
For each optimization, the Pareto optimal front is built (using
the Non-Dominated Sorting Method) considering the set of non-
dominated solutions: thus, if a solution results not to be domi-
nated, at least in one of the two objective functions, it belongs to
the Pareto frontier.
The reference values for the model representing the AOC 15/50
Sandia blade are evaluated in a preliminary structural analysis: the
total deformation at the reference point results d0 = 219.46 mm
and the blade total mass is m0 = 85.78 kg. The BEM analysis of
the original model indicates a generated power P0 of 36.72 kW
for an unperturbed wind speed of 10 m/s, here assumed as the ref-
Fig. 8. Location of the reinforcements (all zones with the exception of 9 and 10, the erence value for the optimization process.
Spar Web and the Spar Flange) inside the original AOC 15/50 blade (from Table 3).
A. Dal Monte et al. / Composite Structures 159 (2017) 144–156 151

Fig. 9. Genetic pool of the original AOC 15/50 blade.

Table 6
Scheme of the crossover function.

Crossover scheme
Parent 1 a b c d e f g h
Parent 2 1 2 3 4 5 6 7 8
Binary vector 1 1 0 0 1 0 0 0
Resulting child a b 3 4 e 6 7 8

Fig. 10. Functional diagram of the S.O.C.R.A.TE. optimization process; the algorithm is based upon the coupling of the commercial code Ansys, an in-house made BEM code
and the commercial code Matlab.

9. First optimization: settings and results The trend of the Pareto Front is shown in Fig. 11. As can be
observed, the solutions identify a clear Pareto Front from the
The main features of the first genetic optimization are summa- 10th generation; subsequent improvements are minimal.
rized in Table 7. The genetic pool of Fig. 12 is referred to one of the optimal indi-
viduals (the 135th of the 27th generation) of the first optimization.
The corresponding values of the fitness functions are: f P = 0.9592
Table 7
and f d = 0.6938. The total measured deformation of 152.26 mm
Settings of the first optimization process. for the reference point shows a great increase in the structural per-
formances (30.62% deformation), while the generated power
First optimization
results of 38.28 kW (+4.25%) at 10 m/s.
Genes 24
Population size 160
Generations 30 10. Second optimization: settings and results
f P upper bound 3
f d upper bound 1.20 In order to focus the optimization on the left side of the Pareto
im penality on f d Quadratic over the 20% Front, a more restrictive upper bound (1.5) is set for the Power fit-
iA penality on f P Quadratic over the 5%
152 A. Dal Monte et al. / Composite Structures 159 (2017) 144–156

Fig. 11. Evolution of the Pareto front for the first optimization process.

Fig. 12. Genetic pool of the 135th individual of the 27th generation (first optimization).

ness function f P . Furthermore, an additional linear penalty function 11. Discussion of the optimization results
is introduced, in order to correct f P when its value exceeds 1.25.
The last generation of the first optimization is adopted as the initial The main geometric features and aerodynamic loads of the orig-
generation of the current optimization process. The main features inal blade and of the (second) optimized one are presented in
of the second genetic optimization are summarized in Table 8. Tables 9 and 10.
The Pareto front for generations No. 1, 5, 10, 15, 20 and 30 is Fig. 15 shows the evolution of the chord distribution for both
shown in Fig. 13. The improvements are less marked than those the original model and the (second) optimized one: chord values
obtained from the previous optimization, however the general are increased along the whole blade span, up to the maximum
trend of the front can be clearly observed. As can be seen, the more value allowed from the constraint imposed on the blade surface.
restrictive bounds on f P force the solutions to assume higher values As a consequence, the blade results to be more loaded in almost
of generated power. every section (see Tables 9 and 10 and Fig. 18): the global amount
Fig. 14 summarizes the genetic pool of the selected solution of aerodynamic forces is 2293 N in the original blade and increases
(corresponding to the 146th individual of the 24th generation) to 2418 N in the (second) optimized one. Moreover, as shown in
for the second optimization. The fitness functions assume the val- Fig. 16, the generalized increment in the twist angle h determines
ues f d = 0.6785 and f P = 0.9574. The resulting displacement of the a corresponding decrease in the angle of attack a for all the consid-
reference point is 148.90 mm (32.15%) and the generated power ered sections (the trend is shown in Fig. 17). In the second part of
is 38.35 kW (+4.44%). the optimized blade, the values of a result to be closer to the angle
that determines the highest aerodynamic efficiency of the adopted
airfoils (6° for the S812 and 5° for the S813). The twist distribution
of AOC 15/50 reveals to be not optimized for the nominal wind
velocity. The blade root has small influence on aerodynamic per-
formance as shown in Fig. 18, however the non-optimized twist
Table 8 of the baseline AOC 15/50 probably causes detachments of the
Settings of the second optimization process. boundary layer and a diffuse stall in the sections near the root.
Second optimization Adopting a higher twist in the root zone helps to reduce these
Genes 24
issues, however it does not greatly affect the overall performances,
Population size 160 due to the small influence the root zone on the power production.
Generations 30 The composite skin layout of the (second) optimized blade is
f P upper bound 1.5 changed by the genetic algorithm. The þ45 lamina (gene A), that
f d upper bound 1.20
covered the entire blade span in the baseline configuration, is
im penality on f d Quadratic over 20%
iA penality on f P Quadratic over 5% removed and replaced by the duplication of the more resistant 0
f P penality on f P Linear over 20% laminas (Genes F, G, I and L). A great improvement in the structural
A. Dal Monte et al. / Composite Structures 159 (2017) 144–156 153

Fig. 13. Evolution of the Pareto front for the second optimization process.

Fig. 14. Genetic pool of the 146th individual of the 24th generation (second optimization).

Table 9
Geometric features and load conditions of the original
AOC 15/50 blade; no angle of attack is computed for
the root section (being characterized by a junction
between an airfoil and an oval section) and for the tip
one (due to Prandtl’s tip loss factor, see [25]).

r [m] c [m] h [°] a [°] F [N]


0.279 0.4570 8.10° – 0
0.775 0.5303 8.10 39.60 19
0.889 0.5462 7.60 36.90 11
1.092 0.5745 6.72 33.86 43
1.702 0.6568 6.24 26.10 38
2.311 0.7173 5.94 19.66 135
3.124 0.7400 5.53 13.39 236
3.937 0.6920 4.86 9.39 301
4.750 0.6258 4.10 8.69 376
5.563 0.5572 3.28 7.55 435
6.603 0.4780 2.37 6.67 366
7.059 0.4410 1.95 5.56 224
7.490 0.4060 1.54 – 109

characteristics is obtained, passing from a 219.46 mm deformation of reminding that the solution of extending the twisted portion of
of the original blade to the 148.90 mm of the (second) optimized the blade up to its root can be find in all commercial Enercon mod-
one (32.15%). els [36]: as is clearly proved in this work, such architecture, besides
In order to better understand the influence of blade geometry increasing the aerodynamic efficiency of the blade portion close to
on its structural behaviour, a further investigation is hereby pro- the nacelle, presents also a not negligible structural benefit.
posed: the (second) optimized blade is analysed using the original The present findings prove that the registered enhancement in
Sandia layout. A total deformation of 187.99 mm is registered the (second) optimized blade are to be ascribed to two
(14.34%), confirming that a structural improvement can also be contributions:
achieved by means of a proper twist distribution. It is just the case
154 A. Dal Monte et al. / Composite Structures 159 (2017) 144–156

Table 10
Geometric features and load conditions of the (second)
optimized AOC 15/50 blade; no angle of attack is
computed for the root section (being characterized by
a junction between an airfoil and an oval section) and
for the tip one (due to Prandtl’s tip loss factor, see
[25]).

r [m] c [m] h [°] a [°] F [N]


0.279 0.4570 20.24 – 0
0.775 0.5766 20.24 28.69 18
0.889 0.5994 19.57 27.56 11
1.092 0.6361 18.41 25.49 40
1.702 0.7179 15.38 19.12 82
2.311 0.7616 12.89 14.23 142
3.124 0.7700 10.24 9.60 249
3.937 0.7338 8.14 7.39 296
4.750 0.6669 6.39 6.46 403
5.563 0.5834 4.85 6.31 466
6.603 0.4723 3.18 6.04 380
7.059 0.4263 2.63 5.27 227
7.490 0.3851 2.29 – 104

 a blade stiffening due to the higher values of both chord and


twist angles along the blade span, responsible for a 14.34%
reduction in the blade deformation
 a more efficient layer distribution, responsible for a 17.81%
reduction in the blade deformation.

Fig. 19 shows the evolution of the flexural rigidity EI along the


blade span, computed as the ratio between the moment M acting
on a given cross section and the rate of rotation dH=dz of the sec-
tion itself (see [10,22]) for the original AOC 15/50 blade, the (sec-
ond) optimized one and the (second) optimized geometry
coupled with the original Sandia layout. A marked improvement
Fig. 15. Comparison between the chord distributions for both the original AOC 15/
50 blade and the (second) optimized one.
can be observed over the whole blade span, particularly in the cen-
tral blade portion.
The surface density distribution along the blade is also pre-
sented in Fig. 20. The overall mass of the blade has been increased,
particularly in the root zone. The two optimized solutions present a
similar trend of the surface density. The results show improve-
ments in the power generation reducing the blade deformation,
however the mass of the blade is increased: a further deeper anal-
ysis should take into account also the Cost Of Energy, not consid-
ered in the present optimization.

12. Conclusion and future work

Both the aerodynamic efficiency and the mechanical character-


istics of the Sandia AOC 15/50 wind turbine blade have been
Fig. 16. Comparison between the twist distributions for both the original AOC 15/
50 blade and the (second) optimized one.
improved using an evolved version of the S.O.C.R.A.TE. algorithm.
For this purpose, the original S.O.C.R.A.TE. formulation has been
modified and, in order to manage a coupled aerodynamic–struc-
tural optimization, a BEM code has been implemented in the algo-
rithm: as a first step, the aerodynamic performance of the blade is
evaluated by the BEM code and the aerodynamic forces acting on
the blade are successively exported for the structural analysis.
The algorithm has succeeded in increasing the power genera-
tion at 10 m/s of an interesting 4.44% by changing both chord
and twist distributions along the blade span. The chord of the opti-
mized blade has been increased in the most part of the sections,
reaching the upper bound of the allowed blade surface. Also the
optimal twist angle distribution has registered a marked increase,
determining an angle of attack perceived by each blade section
close to the one of maximum aerodynamic efficiency. The second
Fig. 17. Comparison between the angle of attack distributions for both the original
part of the S.O.C.R.A.TE. algorithm has lead to an optimized
AOC 15/50 blade and the (second) optimized one.
sequence of the laminas composing the blade layout, determining
a marked decrease of the total deformation at blade tip. Neverthe-
A. Dal Monte et al. / Composite Structures 159 (2017) 144–156 155

Fig. 18. Comparison between the aerodynamic force distributions evaluated in 12 blade sections for both the original AOC 15/50 blade and the (second) optimized one.

[2] Liu S, Janajreh I. Development and application of an improved blade element


momentum method model on horizontal axis wind turbines. Int J Energy
Environ Eng 2012;3.
[3] Refan M, Hangang H. Aerodynamic performance of a small horizontal axis
wind turbine. J Sol Energy Eng 2012;134(2).
[4] ElQatary I, Elhadidi B, Comparison of CFD with BEM technique for wind turbine
simulation of thin and thick rotor blades. In: 31st AIAA applied aerodynamics
conference, June 24–27 2013, San Diego, CA.
[5] Kong C, Bang J, Sugiyama Y. Structural investigation of composite wind turbine
blade considering various load cases and fatigue life. Energy August September
2005;30(11-12,):2101–14.
[6] Mendez J, Greiner D. Wind blade chord and twist angle optimization using
genetic algorithms. In: Proceedings of the fifth international conference on
engineering computational technology. Stirlingshire, UK: Civil-Comp Press;
2006.
[7] Winter G, Periaux J, Galan M, Cuesta P. Genetic algorithms in engineering and
Fig. 19. Comparison of the flexural ridigity distribution for the original AOC 15/50 computer science. Jhon Wiley and Sons; 1995.
[8] Benini E, Toffolo A. Optimal design of horizontal-axis wind turbines using
blade, the (second) optimized one and the (second) optimized geometry coupled
blade-element theory and evolutionary computation. J Sol Energy Eng
with the original Sandia layout.
November 2002;124(4):357–63.
[9] Cai X, Zhu J, Pan P, Gu R. Structural optimization design of horizontal-axis wind
turbine blades using a particle swarm optimization algorithm and finite
element method. Energies Nov 2012;5:4683–96.
[10] Dal Monte A, Raciti Castelli M, Benini E. Multi-objective structural
optimization of a HAWT composite blade. Compos Struct December
2013;106:362–73.
[11] Hu W, Han I, Park SC, Choi DH. Multi-objective structural optimization of a
HAWT composite blade based on ultimate limit state analysis. J Mech Sci
Technol January 2012;26(1):129–35.
[12] Pourrajabiana A, Afsharb PAN, Ahmadizadehc M, Woodd D. Aero-structural
design and optimization of a small wind turbine blade. Optim Methods Renew
Energy Syst Des March 2016;87(Part 2):837–48.
[13] Bottasso CL, Campagnolo F, Croce A. Multi-disciplinary constrained
optimization of wind turbines. Multibody Syst Dyn January 2012;27(1):21–53.
[14] Ashuri T, Zaaijer MB, Martins JRRA, van Bussel GJW, van Kuik GAM.
Multidisciplinary design optimization of offshore wind turbines for
minimum levelized cost of energy. Renew Energy August 2014;68:893–905.
[15] Grujicic M, Arakere G, Pandurangan B, Sellappan V, Vallejo A, Ozen M.
Multidisciplinary design optimization for glass-fiber epoxy-matrix composite
5 MW horizontal-axis wind-turbine blades. J Mater Eng Perform November
Fig. 20. Comparison between the surface density distribution for the original AOC
2010;19(8):1116–27.
15/50 blade and the optimized solutions. [16] Zhu J, Cai X, Gu R. Aerodynamic and structural integrated optimization design
of horizontal-axiswind turbine blades. Energies January 2016;9(66).
[17] Wang YZ, Li F, Zhang X, Zhang WM. Composite wind turbine blade
less, the registered increase in the flexural rigidity of the blade has aerodynamic and structural integrated design optimization based on RBF
meta-model. Mater Sci Forum 2015;813:10–8.
proved to be ascribed also to a blade stiffening due to the higher [18] Ghommema M, Colliera N, Niemib AH, Caloa VM. On the shape optimization of
values of both chord and twist angles along the blade span. flapping wings and their performance analysis. Aerosp Sci Technol January
Further developments could consider different fitness func- 2014;32(1):274–92.
[19] Gillebaart E, De Breuker R. Low-fidelity 2D isogeometric aeroelastic analysis
tions, such as the evaluation of the annual energy production and
and optimization method with application to a morphing airfoil. Comput
the cost of energy. Different materials, like modern GFRPs or CFRPs Methods Appl Mech Eng June 2016;305:512–36.
could also be considered. Finally, a more accurate estimation of the [20] Zingg DW, Nemec M, Pulliam TH. A comparative evaluation of genetic and
gradient-based algorithms applied to aerodynamic optimization. Eur J Comput
blade aerodynamic performance could be obtained by replacing
Mech 2008;17(1-2):103–26.
the BEM code with a CFD solver. [21] Jacobson R, Meadors M, Jacobson E, Link H. Power performance test report for
the AOC 15/50 wind turbine, test B. Colorado: National Wind Technology
References Center, National Renewable Energy Laboratory; 2003.
[22] McKittrick LR, Cairns DS, Mandell J, Combs DC, Rabern DA, Van Luchene RD.
Analysis of a composite blade design for the AOC 15/50 wind turbine using a
[1] Burton T, Jenkins N, Sharpe D, Bossanyi E. Wind energy handbook. John Wiley finite element model, SAND2001-1441; May 2001.
& Sons; 2001. [23] AOC 15/50 Detailed Specification, <http://seaforthenergy.com/wp-content/
uploads/2010/12/AOC1550-Specification-Sheet.pdf>.
[24] Nrel, FAST software, <http://wind.nrel.gov/designcodes/simulators/fast/>.
156 A. Dal Monte et al. / Composite Structures 159 (2017) 144–156

[25] Hansen MOL. Aerodynamics of wind turbines. Earthscan; 2008. [33] Lindenburg C. Investigation into rotor blade aerodynamics – analysis of the
[26] Manwell JF, McGowan JG, Rogers AL. Wind energy explained: theory, design stationary measurements on the UAE Phase-VI rotor in the NASA-AMES wind
and application. John Wiley & Sons Inc.; 2002. tunnel. ECN-C-03-025; July 2003.
[27] Glauert H, Airplane propellers. In: Durand WF, editor. Aerodynamic theory, [34] Yu G, Shen X, Zhu X, Du Z. An insight into the separate flow and stall delay for
vol. 4. Division L. Berlin: Julius Springer. p. 169–360. HAWT. Renew Energy 2011;36:69–76.
[28] Clifton-Smith MJ. Wind turbine blade optimisation with tip loss corrections. [35] Lanzafame R, Messina M. Advanced brake state model and aerodynamic post-
Wind Eng 2009;33:477–96. stall model for horizontal axis wind turbines. Renew Energy 2013;50:415–20.
[29] Lanzafame R, Messina M. Power curve control in micro wind turbine design. [36] Enercon blade models, <http://www.enercon.de/>.
Energy 2010;35:556–61. [37] Benini E, Toffolo A. Optimal design of horizontal-axis wind turbines using
[30] Pratumnopharat P, Leung PS. Validation of various windmill brake state blade-element theory and evolutionary computation. J Sol Eng November
models used by blade element momentum calculation. Renew Energy 2002;124:357–63.
2011;36:3222–7. [38] Wang L, Tanga X, Liu X. Optimized chord and twist angle distributions of wind
[31] Drela M. XFoil: an analysys and design system for low reynolds number turbine blade considering Reynolds number effects. In: International
airfoils, technical report. Cambridge, Massachusetts: MIT Dept. of Aeronautics conference on wind energy: materials, engineering and policies, November
and Astronautics; 1989. 2012.
[32] Timmer WA, van Rooij PRJOM. Summary of the Delft University wind turbine [39] Mendez J, Greiner D. Wind blade chord and twist angle optimization using
dedicated airfoils. J Sol Energy Eng 2003;125(4):488–96. genetic algorithms. In: Fifth international conference on engineering
computational technology. Las Palmas de Gran Canaria (Spain); 2006.

You might also like