We 1605

Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

WIND ENERGY

Wind Energ. (2013)


Published online in Wiley Online Library (wileyonlinelibrary.com). DOI: 10.1002/we.1605

RESEARCH ARTICLE

Vibrational analysis of the flexo-torsional modes of


the NREL 5-MW reference wind turbine
Alejandro D. Otero1,2 , Fernando L. Ponta2 and Lucas I. Lago2
1 CONICET and College of Engineering, University of Buenos Aires, Buenos Aires, Argentina
2 Department of Mechanical Engineering-Engineering Mechanics, Michigan Technological University, Houghton,
Michigan 49931, USA

ABSTRACT
A key for a breakthrough in wind turbine technology is to reduce the uncertainties related to blade dynamics by the
improvement of the quality of numerical simulations of the fluid–structure interaction process. A fundamental step in that
direction is the implementation of structural models capable of capturing the complex features of innovative prototype
blades, so they can be tested at realistic full-scale conditions with a reasonable computational cost.
We make use of a code based on the generalized Timoshenko theory that has the capacity of reducing the geometrical
complexity of the blade section into a stiffness matrix for an equivalent beam, allowing accurate modeling of a three-
dimensional structure of the blade as a one-dimensional finite-element problem. This model is combined with an advanced
flow model based on a re-implementation of the classical blade element momentum theory.
We studied the vibrational modes of the National Renewable Energy Laboratory 5-MW reference wind turbine
blades, which is representative of state-of-the-art multi-megawatt commercial wind turbines. The geometry of the blade
structure was refined by means of a novel interpolation technique matching the properties of the original airfoil sections.
We present results of full-scale simulations of the structural response of these composite laminate blades, which take
into account the effects of their complex internal structure. The fundamental frequencies and vibrational modes are com-
puted for the deformation state of the blade when operating steadily at nominal conditions. We also show results for
stresses on the blade section and for the displacements and rotations of the blade sections along the span. Copyright © 2013
John Wiley & Sons, Ltd.
KEYWORDS
wind turbine; blade structural modeling; generalized Timoshenko model; finite-element method

Correspondence
Fernando L. Ponta, Department of Mechanical Engineering-Engineering Mechanics, Michigan Technological University,
1400 Townsend Drive, Houghton, Michigan 49931, USA.
E-mail: [email protected]

Received 21 June 2012; Revised 20 November 2012; Accepted 21 January 2013

1. INTRODUCTION
Limitations in the current blade technology constitute a technological barrier that needs to be overcome to continue
improvement in wind energy cost. Blade manufacturing is mostly based on composite laminates, which is labor-intensive
and requires highly qualified manpower. It constitutes a bottleneck to turbine upscaling that reflects into the increasing
share of the cost of the rotor, within the total cost of the turbine, as turbine size increases. Figure 1 shows a compilation
of data by the National Renewable Energy Laboratory (NREL), Department of Energy1 on the proportional cost of each
subsystem for different sizes of wind turbines, where the systematic increase of the rotor cost share is clearly reflected.
Moreover, although the rest of the wind turbine subsystems are highly developed technological products, the blades
are unique. There is no other technological application that uses such a device, so practical experience in blade manufac-
turing is relatively new. Blades also operate under a complex combination of fluctuating loads, and huge size differences
complicate extrapolation of experimental data from the wind tunnel to the prototype scale. Hence, computer models of
fluid–structure interaction phenomena are particularly relevant to the design and optimization of wind turbines. The wind
turbine industry is increasingly using computer models for blade structural design and the optimization of its aerodynamics.

Copyright © 2013 John Wiley & Sons, Ltd.


Vibrational analysis of the flexo-torsional modes of the NREL 5-MW RWT A. D. Otero, F. L. Ponta and L. I. Lago

Figure 1. Evolution of the proportional cost for the different wind turbine subsystems, as size increases (data compilation from the
National Renewable Energy Laboratory1 ).

Nevertheless, several features of the complex interaction of physical processes that characterize the coupled aeroelastic
problem still exceed the capacities of existing commercial simulation codes. Changes in structural response, due to the
development of new techniques in blade construction and/or the use of new materials, would also represent a major factor
to take into account if the development of a new prototype blade is considered.
Hence, a key factor for a breakthrough in wind turbine technology is to reduce the uncertainties related to blade dynam-
ics, by the improvement of the quality of numerical simulations of the fluid–structure interaction process and by a better
understanding of the underlying physics. The current state of the art is to solve the aeroelastic equations in a fully non-linear
coupled mode using Bernoulli or Timoshenko beam models (see Hansen et al.,2 where a thorough coverage of the topic is
presented). The goal is to provide the industry with a tool that helps them introduce new technological solutions to improve
the economics of blade design, manufacturing and transport logistics, without compromising reliability. A fundamental
step in that direction is the implementation of structural models capable of capturing the complex features of innovative
prototype blades, so they can be tested at realistic full-scale conditions with a reasonable computational cost. To this end,
we developed a generalized Timoshenko code3 based on a modified implementation of the variational asymptotic beam
sectional technique proposed by Hodges et al. (see Hodges4 and references therein). The ultimate goal is to combine this
code with an advanced non-linear adaptive model of the unsteady flow, based on the vorticity–velocity formulation of the
Navier–Stokes equations, called the kinematic Laplacian equation model,5,6 which would offer performance advantages
over the present fluid–structure solvers.
In this paper, we apply our code to the study of the structural response of the NREL 5-MW reference wind turbine (RWT)
blade as defined by Jonkman et al.7 On the basis of the REpower 5M wind turbine (Hamburg, Germany), the NREL’s RWT
was conceived for both onshore and offshore installations and is well representative of state-of-the-art utility-scale multi-
megawatt commercial wind turbines. The geometry of the blade structure was refined by means of a novel interpolation
technique matching the properties of the original airfoil sections. We present results of full-scale simulations of the struc-
tural response of these composite laminate blades, which take into account the effects of their complex internal structure.
The fundamental frequencies and vibrational modes are computed for the deformation state of the blade when operating
steadily at nominal conditions. We also show results for stresses on the blade section and for the displacements and rotations
of the blade sections along the span.

2. STRUCTURAL MODEL
Wind turbine blades are typically slender structures with high flexibility. They are usually not simple to model because of
the inhomogeneous distribution of material properties and the complexity of their cross sections (Figure 2). Detailed two-
dimensional (2-D) shell or fully three-dimensional (3-D) solid models can be highly expensive, in computational terms, if
the structural problem is required to be solved along many time steps and many different model designs. On the other hand,
the ad hoc kinematic assumptions made in classical theories (such as the Bernoulli or standard Timoshenko approaches)
may introduce significant errors, especially when the blade is vibrating with a wavelength shorter than its length.
Complex blade geometry due to an aerodynamic/mechanical design, new techniques of blade construction and the use
of new materials combine themselves to give a new dimension to the problem.
To obtain a fluid–structure interaction model capable of dealing with the complex features of new-generation blades,
we developed a generalized Timoshenko code3 based on a modified implementation of the variational asymptotic beam
sectional model proposed and developed by Prof. Hodges and his collaborators (see, for example, Hodges4 and references
therein). This model, which is able to work with curved and twisted composite beams, uses the same variables as the
classical Timoshenko beam theory, but the hypothesis of beam sections remaining planar after deformation is abandoned.

Wind Energ. (2013) © 2013 John Wiley & Sons, Ltd.


DOI: 10.1002/we
A. D. Otero, F. L. Ponta and L. I. Lago Vibrational analysis of the flexo-torsional modes of the NREL 5-MW RWT

Figure 2. Example of blade section structural architecture representative of current commercial blade designs. The primary structural
member is a box spar, with a substantial buildup of spar cap material between the webs. The exterior skins and internal shear webs
are both sandwich construction with triaxial fiberglass laminate separated by a balsa core (from Griffin8 ).

Instead, the real warping of the deformed section is interpolated by a 2-D finite-element mesh, and its contribution to the
strain energy is put in terms of the classical one-dimensional (1-D) Timoshenko’s variables by means of a pre-resolution.
The geometrical complexity of the blade section and/or its material inhomogeneousness is reduced into a stiffness matrix
for the 1-D beam. The reduced 1-D strain energy is equivalent to the actual 3-D strain energy in an asymptotic sense.
Elimination of the ad hoc kinematic assumptions in complex geometries produces a fully populated 66 symmetric matrix
for the 1-D beam, with as many as 21 independent stiffnesses, instead of the six fundamental stiffnesses of the original
Timoshenko theory. This is why it is referred to as a generalized Timoshenko theory.
Even for the case of large displacements and rotations of the beam sections, this model allows for accurate modeling
of the bending and transverse shear in two directions, extension and torsion of the blade structure as a 1-D finite-element
problem. Thus, we are able to decouple a general 3-D non-linear anisotropic elasticity problem into a linear 2-D cross-
sectional analysis (which may be solved a priori) and a non-linear 1-D beam analysis for the global problem, which is
what is needed at each time step of a fluid–structure interaction analysis. The cross-sectional 2-D analysis (which may be
performed in parallel for all the cross sections along the blade) calculates the 3-D warping functions asymptotically and
finds the constitutive model for the 1-D non-linear beam analysis of the blade. After one obtains the global deformation
from the 1-D beam analysis, the original 3-D fields (displacements, stresses and strains) can be recovered a posteriori using
the already calculated 3-D warping functions.
In what follows, we briefly outline the generalized Timoshenko theory. Referring to Figure 3, the beam is represented
by a reference line R in the undeformed configuration, which could be twisted and/or curved. At every point along R an
associated orthogonal triad (B1 , B2 , B3 ) is defined in such a way that B1 is tangent to R and B2 and B3 are contained

Figure 3. Generalized Timoshenko theory: schematic of the reference line, orthogonal triads and beam sections before and after
deformation (adapted from Yu et al.9 ).

Wind Energ. (2013) © 2013 John Wiley & Sons, Ltd.


DOI: 10.1002/we
Vibrational analysis of the flexo-torsional modes of the NREL 5-MW RWT A. D. Otero, F. L. Ponta and L. I. Lago

into the section plane, which is normal to R. A correspondent coordinate system .X 1 ; X 2 ; X 3 / is defined, where X 1 is the
coordinate along R and X 2 and X 3 are the Cartesian coordinates on the section plane.
The position of a generic point on each section in the undeformed configuration may be written as

R.X i / D R.X
N 1 / C X ˛ B .X 1 /
˛ (1)

where R N denotes the position of the center of the coordinate system along R, the index ˛ assumes values 2 and 3, and we
make use of the convention that repeated indexes are summed over their rank. The second term in equation (1) represents
the relative position of the point in the cross section. When the structure is deformed because of loading, the original ref-
erence line R adopts a new geometry r, and we have a new triad t1 ; t2 ; t3 associated to each point, where t1 is tangent to
r and t2 and t3 are contained into the normal plane. The material point whose original position was given by R.X i / has,
after the deformation, the position vector

r.X i / D R
N C u C X ˛ t C wi t
˛ i (2)

In equation (2), the vector u is the displacement of the reference line after deformation, the third term represents the posi-
tion of the point in the cross section after deformation considering no shear deformation and the last term is the contribution
to the displacement of the point due to shear and warping of the section. wi are the section warping functions.
The strain energy density is defined as

2U D hh T S ii (3)


R p
where S is the matrix of material characteristics expressed in the Bi coordinates, hhii D s  G dX 2 dX 3 , s is
the domain of a cross section and G is the determinant of the metric in the undeformed base. In this case, ij D
.1=2/.Fij C Fj i /  ıij is the Jaumann–Biot–Cauchy strain tensor, Fij are the components of the gradient-of-deformation
tensor defined as Fij D ti  gk Gk  Bj , with gk and Gk as the covariant base vectors for the deformed configuration
and the contravariant base vectors in the undeformed configuration, respectively, obtained from the kinematic description
of equations (1) and (2).  provides a suitable measure of the 3-D strain field in terms of the beam strain measures and
arbitrary warping functions.
The next step is to find a strain energy expression in terms of the deformation measures of the equivalent 1-D
beam, which asymptotically matches the 3-D strain energy. This is achieved by using the variational asymptotic method
proposed by Berdichevsky10 to find the relation between beam section warping and deformation measures that minimize
the energy density. Finally, the energy density is expressed in terms of 1-D deformation measures equivalent to those of the
Timoshenko theory. For details, see Otero and Ponta.3
This minimization process is carried on by the solution of a 2-D finite-element problem on each blade section. For the
discretization of the 2-D sections, we adopted a tri-quadrilateral finite-element technique (for details, see Ponta6 ), which is
based on the use of nine-node biquadratic isoparametric finite elements that possess a high convergence rate and the ability
of reducing the so-called skin error on curvilinear boundaries.11 The advantages of the nine-node quadrilateral isoparamet-
ric element are combined in the tri-quadrilateral technique with the geometrical ability of triangular grids to create suitable
non-structured meshes with gradual and smooth changes of mesh density.
Finally, a stiffness matrix for the 1-D beam problem SN is formed in such a way as to obtain a functional for the strain
energy density of expression (3) of the form

2U D N T SN N (4)
 
  
where N D is the array of Timoshenko measures of deformation, with  T D 11 212 213 the extension

 
and transverse shear strains and T D 1 2 3 the strain measures due to torsion and bending.
To solve the 1-D problem for the equivalent beam, we use a formulation based on the intrinsic equations of the beam
obtained from variational principles12 and weighted in an energy-consistent way according to Patil and Althoff,13 which
produces the following variational formulation:
2 3 2
Z ` Z `
4ı VN T N PN C ı FN T SN 1 FPN 5 dX 1 D 6 N T N0 NT K N T O FN
O FN C ı V
„ ƒ‚I V … „ ƒ‚ … 4„ı Vƒ‚F… C „
ıV ƒ‚ …
0 0 „ ƒ‚ …
1 2 3 4 5
3

N T Nf  ı V
C „ƒ‚…
ıV NT V N 0 ı FN T K
N ı FN T V
O IN V N7
O T N  ı FN T O T V 1
„ ƒ‚ …C„ ƒ‚ …  „ ƒ‚ V … „ ƒ‚ …5 dX (5)
6 7 8 9 10

Wind Energ. (2013) © 2013 John Wiley & Sons, Ltd.


DOI: 10.1002/we
A. D. Otero, F. L. Ponta and L. I. Lago Vibrational analysis of the flexo-torsional modes of the NREL 5-MW RWT

where
     
F V f
FN D ; N D
V ; Nf D
M  m

     
Q 0 Q
 0 KQ 0
O D ; O D
V ; O D
K
Q Q Q
V Q
 eQ1 Q
K
A tilde indicates the skew-symmetric matrix associated to a vector magnitude in such a way that, for example, if we
have any pair of vectors A and B, the matrix–vector product A Q B is equivalent to the cross product A  B. Thus, Q is
associated with , Q with , V Q with V and so forth. Hence, matrix O is a re-arrangement of the components of the strain
measure vector N defined earlier, the generalized velocity vector V N and matrix V
O represent the components of the linear and
angular velocities and matrix KO represents the initial torsion and curvatures of the beam (matrix eQ1 is the skew-symmetric
 
matrix associated to eT1 D 1 0 0 , the unit vector along X 1 ). The generalized force vector FN represents the forces
and moments related to the strain measures (N D SN 1 F), N and the generalized distributed load vector Nf represents the forces
and moments distributed along the axis of the beam. Here, SN is the stiffness matrix for the 1-D model corresponding to
equation (4); and IN is the inertia matrix of each section. The upper dot indicates a time derivative, and the prime a derivative
with respect to the longitudinal coordinate of the beam X 1 .
This variational formulation was discretized by the spectral element method.14,15 The magnitudes in (5) were replaced
by their interpolated counterparts, V N D He Qe and FN D He Qe , where He and He are the interpolation function arrays
N
V FN VN FN
e
and Q is a vector containing the nodal values of both the generalized velocities and the generalized forces. Superscript ‘e’
indicates discretization of the terms at the elemental level, which will disappear after the final assembly of the terms into the
global matrix for the whole beam. The axial derivatives of the magnitudes were interpolated in a similar way, V N 0 D Be Qe
N
V
and FN 0 D Be Qe , where Be and Be are the arrays for the interpolation function derivatives. Then, the discretized version
FN N
V FN
of (5) is obtained,
 
ıQe T Me1 Q
P e D ıQe T Ke C Ke Qe C ıQe T Ke qN e C ıQe T Be .Qe /
1 2 q Q (6)

where
Z 1 h i
Me1 D HeVN T IN HeVN C HeFN T SN 1 HeFN J dt
1
Z 1 h i
Ke1 D HeVN T BeFN C HeFN T BeVN J dt Qe
1
Z 1 h i
Ke2 D HeVN T K
O He  H e T K
NF NF
O T He J dt
N
V
1
Z 1
Keq D HeVN T HeFN J dt
1

Me1 corresponds to the discretization of terms 1 and 2, giving the equivalent of a mass matrix. Ke1 , corresponding to terms 3
and 8, is the stiffness matrix of the 1-D problem. Ke2 , corresponding to terms 4 and 9, is the additional stiffness related with
the twist and curvature of the undeformed configuration. Keq corresponds to the evaluation of term 6, the contribution of
the distributed loads; and qN e is an array containing the nodal values of the generalized distributed loads. t is the natural
coordinate in the elements, and J is the Jacobian of the mapping from the problem coordinate X 1 to t.11 The discretized
version of the terms in (5) related to non-linear interactions, i.e., terms 5, 7 and 10, gives
Z 1h i
BeQ .Qe / D HeVN T O HeFN  HeVN T V
O IN He  He T O T He Qe J dt
N
V N
F N
V
1

A linearization of BeQ .Qe / around any given configuration Qe1 gives the matrix
Z 1n h i
 
KeN Qe1 D O 1 IN He  FO 1 SN 1 He C PO 1 He
HeVN T O 1 HeFN  V N
V FN N
V
1
h io
e T O T N 1 e T e
CH V S H  O 1 H
FN 1 FN
J dt N
V

where
   
0 FQ 0 PQ v
FO D ; PO D
FQ MQ PQ v PQ !

Wind Energ. (2013) © 2013 John Wiley & Sons, Ltd.


DOI: 10.1002/we
Vibrational analysis of the flexo-torsional modes of the NREL 5-MW RWT A. D. Otero, F. L. Ponta and L. I. Lago

Matrix FO is a re-arrangement of the components of the generalized forcevectorFN defined earlier. Matrix PO is a
Pv
re-arrangement of the components of the generalized momentum vector PN D , which represents the linear and
P!
angular momenta related with the generalized velocities (PN D IN V).
N The tilde operates in the same way defined before, and
the subscript 1 indicates the value of the magnitudes at a given state Qe1 .
Finally, after the assembly of the elemental terms into the global system, the solution for the non-linear problem (5) in
its steady state was obtained by solving iteratively for Q the discretized expression
h  i 
K1 C K2 C KN Q.i/ Q D Kq qN  .K1 C K2 /Q.i/  BQ Q.i/ (7)

and updating the global vector of nodal values of the generalized velocities and forces as Q.iC1/ D Q.i/ C Q.
From the steady-state solution, we also obtain the vibrational modes of the blade structure and their corresponding
frequencies by solving the eigenvalue problem
h  i
P C K1 C K2 C KN Q.i/ Q D 0
M1 Q (8)

From these results for the intrinsic equations, we recovered the displacements and rotations of the blade sections by
solving the kinematic equations for the beam4

u0  CTrR . C e1 / C e1 C KQ uD0 (9)


Q C Q C C0 CT  CrR K
K Q CT D 0 (10)
rR rR rR

where u is the vector of displacements of each point along the reference line from its position in the reference config-
uration to the one in the deformed configuration and CrR is the orthogonal matrix that rotates the local triad from its
original orientation in the reference configuration to the one in the deformed configuration (both are defined in the function
of the longitudinal coordinate X 1 ). The strains  and  were computed from the generalized forces and stiffness of the
corresponding blade section. Equations (9) and (10) were also linearized and, like the other expressions, discretized by the
spectral element method.

3. AERODYNAMIC MODEL
The flow model that interacts with the structural counterpart presented in Section 2, called large sectional rotation blade
element momentum (BEM), is responsible for providing the aerodynamic loads along the rotor blades and is sensitive
enough to take into account all the complex deformation modes that the structural model is able to solve. The basis for our
aerodynamic model is the well-known BEM. Nevertheless, because of the high level of detail that our structural model can
provide, a complete re-formulation was needed in the aerodynamic model to obtain a compatible level of description.
The tendency in the wind turbine industry to increase the size of the state-of-the-art machine7 drives the development
of not only bigger but also more flexible blades that are lighter. It is observed for this type of wind turbine blades that
big deformations, due to either blade flexibility or pre-conforming processes, produce high rotations of the blade sections.
Moreover, blades could be pre-conformed with specific curvatures given to any of their axes (i.e., coning/sweeping). This
tendency puts in evidence one of the most important limitations of the current BEM theory. Although the basics of this
theory keeps being perfectly valid, the actual mathematical formulation implies the assumption of blade sections remaining
perpendicular to an outwards radial line contained in the plane of the actuator disk coincident with the rotor’s plane. That
is, even though the basics of the BEM theory (i.e., the equation of the aerodynamic loads on a set of blade elements and the
change of momentum in a set of concentric streamtubes) keeps being valid, the mathematical formulation cannot represent
large rotations of the blade sections. This basically leads to a misrepresentation of the effects of the large deformation asso-
ciated to flexible blades on the computation of the aerodynamic loads. Hence, a new mathematical formulation is required
to project the velocities obtained from the momentum theory onto the blade element’s plane and then re-project backwards
the resulting forces from blade element theory onto the plane of the streamtube actuator disk.
In what follows, we will describe the main characteristics of our model and refer to Manwell et al.16 and Burton et al.17
for details on the classical BEM theory. We start by defining a set of orthogonal matrices that perform the rotation of the
physical magnitudes involved (velocities, forces, etc.) The interaction with external control modules will require a con-
stant update of this projection matrices. For example, the rotor azimuth matrix, besides the instantaneous position of the
blade along its rotation, can reflect control actions on the dynamics of the electromechanical train that define the rotor’s
angular speed .

Wind Energ. (2013) © 2013 John Wiley & Sons, Ltd.


DOI: 10.1002/we
A. D. Otero, F. L. Ponta and L. I. Lago Vibrational analysis of the flexo-torsional modes of the NREL 5-MW RWT

The components of the wind velocity vector facing the differential annulus of our actuator disk Wh are affected,
according to BEM theory, by the axial induction factor a and the rotational induction factor a0 .

2 3
Wwh.1/ .1  a/
6 7
Wh D 4 Wwh.2/ C  rh .1 C a0 / 5 (11)
Wwh.3/

where Wwh is the incoming wind velocity projected into the h coordinate system,  is the angular velocity of the rotor and
rh is the radial distance of the airfoil section in the h coordinate system. The h subscript here indicates that the wind velocity
vector is described in the hub coordinate system (Figure 4) according to standards from the International Electrotechnical
Commission.18
Then, to compute the relative velocity affecting a blade element, we will project Wh going through the different coordi-
nate systems, from the hub until the blade’s section coordinate system. This includes orienting the wind velocity for coning
effects on the rotor and any existing pitch angle defined for the blade. Thus, the coning rotation matrix Ccn is a linear
operator with a basic rotation taking into account the coning angle for the rotor (Figure 5).
The pitching rotation matrix, Cp , represents a rotation around the z-axis resulting from the previous coning rotation
according to the pitch of the blade, giving rise to the blade’s system of coordinates referred to by the ‘b’ subscript (Figure 6).

Figure 4. Hub coordinate system according to standards from the International Electrotechnical Commission.18

Figure 5. Cone and tilt angles for upwind wind turbines, according to standards from the International Electrotechnical Commission.18

Wind Energ. (2013) © 2013 John Wiley & Sons, Ltd.


DOI: 10.1002/we
Vibrational analysis of the flexo-torsional modes of the NREL 5-MW RWT A. D. Otero, F. L. Ponta and L. I. Lago

Figure 6. Blade coordinate system according to standards from the International Electrotechnical Commission.18

For example,
2 3
cos.p /  sin.p / 0
Cp D 4 sin.p / cos.p / 0 5 (12)
0 0 1

where p is the pitch angle.


Two more re-orientations are needed to obtain the instantaneous coordinate system of the blade sections, associated
with the deformed reference line r of Section 2. The first of this matrices contains information on the blade section’s
geometry at the time the blade was designed and manufactured. As it was mentioned previously, the blade could have pre-
conformed curvatures along its longitudinal axis (i.e., the blade axis is no longer rectilinear). These curvatures can reflect
either an initial twist along the longitudinal axis or a combination of twist plus pre-bending on the other two axes (i.e.,
coning/sweeping). To this end, we compute during the blade design stage a set of transformation matrices that contain the
information of the 3-D orientation of the blade’s sections for each position on the longitudinal axis, R in Figure 3, as we
move along the span. We call the corresponding matrix of each section CRb , as it relates the global coordinate system of
the blade b, with the system of coordinates of the blade sections in the undeformed configuration defined by line R, as in
Section 2.
After the CRb is applied, one more projection is needed to obtain the instantaneous coordinate system associated with r.
This last transformation is given by the CrR matrix, computed by the 1D structural model (equation (10)). It contains infor-
mation to transform vectors from the R to the r systems after structural deformations had occurred. Note that this matrix
is updated at every time step of the 1D model during dynamic simulations, being one of the key variables transporting
information between the structural and aerodynamic models.
After all these projections of the Wh vector, we have the relative wind velocity expressed in the blade’s section
coordinate system. The expression for the flow velocity relative to the blade section is

Wrel D .CrR CRb Cp Ccn Wh / C vstr (13)

where the addition of vstr corresponds to the blade section structural deformation velocities, coming from the structural
model.
Then, the magnitude jWrel j and the angle of attack ˛ are used to compute the forces on the airfoil section through the
aerodynamic coefficients Cl and Cd . Another innovation of our model is that the data tables from a static wind tunnel are
corrected at each time step to consider either rotational-augmentation or dynamic-stall effects or both.
The aerodynamic loads acting on the blade element are then projected back onto the h coordinate system,

dFh D CTcn CTp CTRb CTrR CLthal dFr (14)

where CLthal is the matrix that projects the lift and drag forces onto the chord-wise and chord-normal directions, which are
aligned with the coordinates of r. Finally, as in the classical BEM theory, dFh is equated to the rate of change of momentum
in the annular streamtube corresponding to the blade element. The component normal to the rotor’s disk is equated to the
change in axial momentum, whereas the tangential component is equated to the change of angular momentum.

Wind Energ. (2013) © 2013 John Wiley & Sons, Ltd.


DOI: 10.1002/we
A. D. Otero, F. L. Ponta and L. I. Lago Vibrational analysis of the flexo-torsional modes of the NREL 5-MW RWT

To apply this theory to HAWT rotors, we must introduce some corrective factors into the calculation process. The BEM
theory does not account for the influence of vortices being shed from the blade tips into the wake on the induced velocity
field. These tip vortices create multiple helical structures in the wake, which play a major role in the induced velocity
distribution at the rotor. To compensate for this deficiency in BEM theory, a tip-loss model originally developed by Prandtl
is implemented as a correction factor to the induced velocity field.19 In the same way, a hub-loss model serves to correct the
induced velocity resulting from a vortex being shed near the hub of the rotor.16,17 Another modification needed in the BEM
theory is the one developed by Glauert20 to correct the rotor thrust coefficient in the ‘turbulent-wake’ state. This correction
plays a key role when the turbine operates at high tip speed ratios, and the induction factor is greater than about 0.45.
The BEM theory was originally conceived for axisymmetric flow. Often, however, wind turbines operate at yaw angles
relative to the incoming wind, which produces a skewed wake behind the rotor. For this reason, the BEM model needs
also to be corrected to account for this skewed wake effect.21,22 The influence of the wind turbine tower on the blade
aerodynamics must also be modeled. We implemented the models developed by Bak et al.23 and Powles,24 which provide
the influence of the tower on the local velocity field at all points around the tower. This model contemplates increases in
wind speed around the sides of the tower and the cross-stream velocity component in the tower near flow field.
Our model also incorporates the possibility to add multiple data tables for the different airfoils and use them in real time
according to the instantaneous aerodynamic situations on the rotor. It also uses Viterna’s extrapolation method25 to ensure
the data availability for a range of angles of attack of ˙180ı .

4. NUMERICAL EXPERIMENTATION
In this section, we report some recent results of the application of our model to the analysis of a set of rotor blades based
on the 5-MW RWT proposed by NREL.7 On the basis of the REpower 5-MW wind turbine, NREL’s RWT was conceived
for both onshore and offshore installations and is well representative of typical utility-scale multi-megawatt commercial
wind turbines. Although full specific technical data are not available for the REpower 5-MW rotor blades, a baseline
from a prototype blade was originally released by LM Glassfiber in 2001 for the Dutch Offshore Wind Energy Converter
(DOWEC) 6-MW wind turbine project26,27 and later re-adapted by NREL. In addition, the NREL 5-MW RWT project has
been adopted as a reference model by the integrated European Union UpWind research program28 and the International
Energy Agency Wind Annex XXIII Subtask 2 Offshore Code Comparison Collaboration.29–31
As stated in NREL’s RWT project, the length of our rotor blade is set to 61.5 m. All basic aerodynamic properties, such
as blade section chords, twist angles and basic spanwise station distribution, correspond to the original data.7 Airfoil types
were also matched from the original blade. Then, an innovative 3D morphing technique based on variational cubic spline
interpolation allows us to obtain smooth transitions along the span of the blade between the airfoil sections of the original
design. This way, we can divide the blade into any number of sections larger than the original and generate many finite-
element meshes for a more refined study of the structural features. After the internal regions and materials were defined,
a tri-quadrilateral mesh was generated for a number of blade sections along the span. As an example, Figure 7 shows the
finite-element meshes of two morphed airfoil sections located at the 20% and 60% of the blade span. Using the technique

0.8
0.6
0.4
0.2
y [m] 0
0.2
0.4
0.6
0.8
2.5 2 1.5 1 0.5 0 0.5 1 1.5
(a)
0.4

0.2
y [m]
0
0.2
2 1.5 1 0.5 0 0.5 1
x [m]
(b)

Figure 7. Finite-element meshes for morphed sections corresponding to (a) 20% of the blade span and (b) 60% of the blade span.

Wind Energ. (2013) © 2013 John Wiley & Sons, Ltd.


DOI: 10.1002/we
Vibrational analysis of the flexo-torsional modes of the NREL 5-MW RWT A. D. Otero, F. L. Ponta and L. I. Lago

described for the internal blade structure components, we refined 46 blade sections along the span to match the structural
properties of the ones reported by NREL.7 The main targeted properties to refine were edgewise, flap-wise and torsional
stiffnesses as well as mass density for every blade section. The pitch axis centering and the location of the aerodynamic
coefficients reference points were also computed according to the information of Jonkman et al.7
The blade structure is a combination of two external aerodynamic shells, mounted on a box-beam spar, which provides
the main structural component to the aerodynamic forces. Analyzing a blade section (Figure 2), we can see the aerodynamic
shells plus two spar caps, which together with two shear webs, form the box-beam spar. Constructive characteristics,
such as thickness as well as the number and orientation of fiberglass layers for the different structural components of the
blade sections, are covered in detail in reports published by the SANDIA National Labs.8,32 These reports also provide a
comprehensive description of lamination sequences and material properties.
According to NREL’s report, the rotor has an upwind orientation and is composed of three blades. The hub diameter is
3 m and is located at 90 m from the ground level. Total rotor diameter is 126 m. It has a pre-cone of 2:5ı and an overhang
distance of 5 m from the tower axis. The rated wind speed for this turbine is 11.4 m s 1, and the corresponding tip speed
ratio is 7.
After stiffness and inertia matrices are computed for a discrete number of cross sections along the span of the blade as
described in Section 2, the calculation of the aeroelastic steady state of the NREL’s RWT blades working under nominal
conditions was solved by fully coupling the structural and aerodynamic models presented in Sections 2 and 3. Boundary
conditions for the beam in this steady-state calculation are defined as follows. In the dynamic equation (7), boundary condi-
tions need to be defined for the six linear and angular velocities and for the six generalized forces, i.e., forces and moments.
We impose null force and moment conditions at the blade tip. At the blade root, we impose rotating clamped conditions,
the peripheral velocity and rotor angular velocity. In the solution of the linear and angular kinematical equations (9) and
(10), zero displacement and rotations are imposed at the blade root, as they are expressed in the blade’s coordinate system.
Figure 8 shows the displacement of the blade’s reference line (blade axis) Uh when it is subjected to the aerodynamic
steady load in normal operational conditions. Figure 9 shows the corresponding rotations of the blade sections  h . These
geometrical magnitudes were referred to as a coordinate system, h from hub, aligned with the rotor’s plane, according to
standards from the International Electrotechnical Commission.18 Hence, the first unit vector is normal to the rotor’s plane
(i.e., axial) pointing downwind, the second is in the rotor’s tangential direction pointing to the blade’s trailing edge and the
third unit vector is in the radial direction pointing to the blade tip.
From Figure 8, we can see that the displacement normal to the rotor’s plane Uh1 of the blade tip is 5.73 m. This is
perfectly consistent with results shown by Jonkman et al.7 Added to this, the tangential displacement Uh2 is 0.78 m in the
negative direction, meaning that aerodynamic forces are bending the blade in the direction towards its rotation as the rotor
is producing a positive driving torque.
In Figure 9, angles h2 and h1 are directly associated with blade bending in the normal and tangential directions to the
rotor plane, which correspond to displacements Uh1 and Uh2 , respectively. It is important to note that angle h2 represents
the angular displacements, which takes the blade’s axis out of the rotor’s plane.

6
Uh1 [m]

0
0 10 20 30 40 50 60 70

0
Uh2 [m]

0.2
0.4
0.6
0.8
0 10 20 30 40 50 60 70

0
Uh3 [m]

0.2
0.4
0.6
0 10 20 30 40 50 60 70
r [m]

Figure 8. Linear displacements of the reference line Uh when the beam is subjected to a steady load in normal operational conditions
(referred to as a coordinate system aligned with the rotor’s plane).

Wind Energ. (2013) © 2013 John Wiley & Sons, Ltd.


DOI: 10.1002/we
A. D. Otero, F. L. Ponta and L. I. Lago Vibrational analysis of the flexo-torsional modes of the NREL 5-MW RWT

1.5

h1 [deg]
1

0.5

0
0 10 20 30 40 50 60 70

15

h2 [deg]
10

0
0 10 20 30 40 50 60 70

0
h3 [deg]

2
0 10 20 30 40 50 60 70
r [m]

Figure 9. Rotations of the beam sections  h when the beam is subjected to a steady load in normal operational conditions (referred
to as a coordinate system aligned with the rotor’s plane).

Next, we solved the eigenvalue problem in expression (8) to obtain the linear modes of vibration around the steady-state
configuration computed before, and their corresponding frequencies. Linear mode analysis provides relevant information
both for the natural blade frequencies, around a steady-state condition, and for the ways the blade deforms in each mode.
Table I summarizes the first 10 modes obtained showing the frequencies together with the corresponding dominant dis-
placements and rotations. Table II shows a comparison for the first three modes reported by NREL in Jonkman et al.7
using FAST33 and ADAMS34 software. FAST and ADAMS are considered today as state-of-the-art software for structural
blade analysis. From this comparison, we see that the frequencies computed with our model match previous studies with a
difference of 1% for the first mode and a maximum difference of 5% for the second and third modes. This difference is not

Table I. List of frequencies for the first 10 modes of vibration in the unloaded case and the aeroelastic
steady-state solution.
Mode Unloaded Aeroelastic steady state

Frequency (Hz) Frequency (Hz) Dominant U Dominant 

1 0.6511 0.7066 Uh1 h2


2 1.0505 1.0188 Uh2 h1
3 1.7570 1.8175 Uh1 h2
4 3.5755 3.3403 Uh2 h3
5 3.9388 3.9493 Uh1 h2
6 6.1181 6.4682 Uh2 h3
7 6.4378 6.6851 Uh1 h3
8 8.0013 8.0129 Uh2 h3
9 8.0514 8.2403 Uh1 h2
10 9.5403 9.7819 Uh1 h2
Also, the dominant components of Uh and  h for the second case are presented.

Table II. Frequency comparison for the first three modes according
to the National Renewable Energy Laboratory report.7
Mode Frequency (Hz) FAST ADAMS

1 0.7066 0.6993 0.7019


2 1.0188 1.0793 1.0740
3 1.8175 1.9223 1.8558

Wind Energ. (2013) © 2013 John Wiley & Sons, Ltd.


DOI: 10.1002/we
Vibrational analysis of the flexo-torsional modes of the NREL 5-MW RWT A. D. Otero, F. L. Ponta and L. I. Lago

0.5

Mode # 1
0

−0.5

−1

0.5

Mode # 4
0

−0.5

−1

1
Mode # 10

0.5

−0.5

−1
0 5 10 15 20 25 30 35 40 45 50 55 60

Figure 10. Amplitude of Uh and  h for three vibrational modes around the aeroelastic steady-state configuration (normalized by the
dominant component). From top to bottom modes # 1, 4 and 10.

Figure 11. The first three components of the Jaumann–Biot–Cauchy stress tensor Z D S  for the section located at 40% of the
blade span (referred to as the undeformed coordinate system .X 1 ; X 2 ; X 3 / in Pa). From top to bottom: Z11 , Z12 and Z13 .

Wind Energ. (2013) © 2013 John Wiley & Sons, Ltd.


DOI: 10.1002/we
A. D. Otero, F. L. Ponta and L. I. Lago Vibrational analysis of the flexo-torsional modes of the NREL 5-MW RWT

surprising as the level of detail and richness of information that our computational tools can register are not present in the
previous software such as FAST or ADAMS.
Figure 10 shows the amplitude of the deformation along the span for the three components of Uh and  h , normalized
by the dominant component, for the 1st, 4th and 10th modes.
After computing the global deformation from the 1-D beam analysis, we recovered the corresponding 3-D fields
(displacements, strains and stresses) using the 3-D warping functions previously calculated with our model. The knowledge
of the stress state of every layer is of utter importance in the analysis of wind turbine blades to improve lifetime and reli-
ability of the design. Our model can provide the full six tensorial components of the stress tensor. Besides, it can provide
the three components of the displacement and six components of the strain.
For the aeroelastic steady state shown previously, we present in Figure 11 the first three components of the Jaumann–
Biot–Cauchy stress tensor Z D S  for the section located at 40% of the blade span. This region of the span is particularly
interesting as it combines in its function a relevant contribution to both energy production and structural support, and it
shows significant stress accumulation compared with other regions along the blade span. The dominant stress component
Z11 is the one primarily associated with the flap-wise bending loads out of the rotor plane, and the shear stress component
in the flap-wise direction Z13 is also consistent with this main-load pattern. Note in the plot of Z11 how the lower spar
cap is subjected to tensile stress whereas the upper one is under compression stress. The shear stress component in the
chord-wise direction Z12 is significantly smaller than its flap-wise counterpart and is consistent with a load pattern indicat-
ing the shear force pointing towards the leading edge, which is responsible for the contribution of the blade to the turbine
driving torque.

5. CONCLUSIONS AND OUTLOOK FOR FURTHER WORK


With the method presented in this paper, we are able to model the structural behavior of wind turbine blades with the
simplicity and economy of a 1-D model but with a level of description equivalent to a much more costly 3-D approach.
The 1-D model is used to compute a fast, but accurate, solution for the deformed state of the blade when subjected to a
steady load in normal operational conditions and an analysis of the vibrational modes around this steady configuration.
This provides a valuable tool to use during the design process. In that sense, the capacity of the generalized Timoshenko
theory to capture the bending–twisting coupled modes in its fully populated 66 stiffness matrix for the 1-D beam problem
plays a fundamental role.
All the deformation modes of the blade are combined modes due to the geometrical complexity and material inhomo-
geneousness in the section, i.e., there are no pure flexural or pure torsional modes. Plots of the vibrational modes may
serve to identify eventual unstable states in the dynamic behavior of innovative prototype blades. Figure 10 shows that, for
certain modes, in some portions of the span, bending due to lift force occurs simultaneously with twisting in the sense that
increases the angle of attack and then the lift force. A complete dynamic analysis of the fluid–structure interaction process
would be needed to determine if those particular modes would be activated or not during the blade operation. Nevertheless,
having the possibility of quickly identifying those modes (and their associated frequencies) at an early stage of the design
process seems very helpful.
Regarding the linear vibrational modes depicted in Figure 10, the first mode shows mainly flap-wise curvatures corre-
sponding to the fundamental frequency of the blade, but as the blade is initially twisted and has complex inhomogeneous
sections, it is not a pure bending mode. Therefore, as a result of the non-conventional couplings, curvature deformation in
the rotor plane and also torsion appears in this mode. The fourth mode is mainly commanded by flexo-torsional coupling;
hence, it could be responsible for fluttering of the blade in case it is excited by the interaction with the surrounding fluid. In
that regard, as it was mentioned earlier, a complete dynamic analysis would be needed to determine which particular modes
would be activated during the blade operation. But the presence of flexo-torsional interaction (such as the one observed in
the fourth mode) could be a good indicator of how good a ‘candidate’ for flutter a mode could be by showing its potential
for bend–twist coupling. The 10th mode is also principally a flap-wise curvature mode but with higher frequency than the
first. It shows a more complex behavior arising from complicated couplings among different deformations. The frequencies
of the first three modes presented in Table II are in good agreement with published results obtained by other models.
The aforementioned flexo-torsional characteristic also gives this model the ability to simulate the dynamic perfor-
mance of adaptive blades, at an affordable computational cost. In the adaptive blade concept (see Griffin35 and Locke36
among others), tailoring of the flexo-torsional modes of the blade is used to reduce aerodynamic loads by controlling the
coupling between bending and twisting. As the blade bends under the load, the angle of attack of the airfoil sections
changes, reducing the lift force. By limiting extreme loads and improving fatigue performance, this added passive control
reduces the intensity of the actuation of the active control system. Plots in Figures 8 and 9 provide valuable information
about the simultaneous deformation of twisting and bending under a given load.
Recovering of the stress tensor components for the different zones of the blade section helps in the prediction of stress
concentration in the basic design, which may ultimately lead to eventual material failure. More exhaustive fatigue studies

Wind Energ. (2013) © 2013 John Wiley & Sons, Ltd.


DOI: 10.1002/we
Vibrational analysis of the flexo-torsional modes of the NREL 5-MW RWT A. D. Otero, F. L. Ponta and L. I. Lago

can be conducted by analyzing the stress in both the steady state or time-marching solutions of the problem. The capability
of computing the whole six components of the stress tensor makes it possible to apply sophisticated failure theories.
We plan to continue our work with a dynamic simulation of the fluid–structure problem. In a first stage, we will couple
the phenomena by feeding back changes in geometry due to blade deformation in our aerodynamic model and re-computing
the forces. At this stage, we also plan to include statistically generated perturbations to represent fluctuations in wind speed
and direction based on anemometry data for wind resource in several representative locations. Besides providing us with
a fast model for a quick analysis, this model will serve as an intermediate step before the ultimate goal of coupling the
structural model with the aforementioned velocity–vorticity kinematic Laplacian equation approach.

ACKNOWLEDGEMENTS
The authors are very grateful for the financial support made available by the National Science Foundation through grants
CEBET-0933058 and CEBET-0952218 and by the University of Buenos Aires through grant 20020100100536 UBACyT
2011/14.

REFERENCES
1. NREL. Wind power today. Report DOE/GO-102005-2115, US Department of Energy, 2005.
2. Hansen MOL, Sørensen JN, Vousitas S, Sørensen N, Madsen HA. State of the art in wind turbine aerodynamics and
aeroelasticity. Progress in Aerospace Sciences 2006; 42: 285–330.
3. Otero AD, Ponta FL. Structural analysis of wind-turbine blades by a generalized Timoshenko beam model. Journal of
Solar Energy Engineering 2010; 132: 011–015.
4. Hodges DH. Nonlinear Composite Beam Theory. AIAA: Reston, Virginia, 2006.
5. Ponta FL. The kinematic Laplacian equation method. Journal of Computational Physics 2005; 207: 405–426.
6. Ponta FL. The KLE method: a velocity–vorticity formulation for the Navier–Stokes equations. Journal of Applied
Mechanics 2006; 73: 1031–1038.
7. Jonkman J, Butterfield S, Musial W, Scott G. Definition of a 5-MW reference wind turbine for offshore system
development. Technical Report NREL/TP-500-38060, National Renewable Energy Laboratory, 2009.
8. Griffin DA. Blade system design studies volume I: composite technologies for large wind turbine blades. Report
SAND2002-1879, Sandia National Laboratories, 2002.
9. Yu W, Hodges DH, Volovoi V, Cesnik CES. On Timoshenko-like modeling of initially curved and twisted composite
beams. International Journal of Solids and Structures 2002; 39: 5101–5121.
10. Berdichevsky VL. Variational-asymptotic method of constructing a theory of shells. Journal of Applied Mathematics
and Mechanics 1979; 43: 664–687.
11. Bathe KJ. Finite Element Procedures. Prentice Hall: Englewood Cliffs, NJ, 1996.
12. Hodges DH. Geometrically exact, intrinsic theory for dynamics of curved and twisted anisotropic beams. AIAA Journal
2003; 41: 1131–1137.
13. Patil MJ, Althoff M. Energy-consistent, Galerkin approach for the nonlinear dynamics of beams using mixed, intrinsic
equations, AIAA/ASME/ASCE/AHE/ASC Structures, Structural Dynamics and Material Conference, AIAA, Reston,
VA, 2006; 1–9.
14. Karniadakis GE, Bullister ET, Patera A. A spectral element method for solution of two- and three-dimensional
time-dependent incompressible navier-stokes equations. In Finite Element Methods for Nonlinear Problems. Springer-
Verlag: New York/Berlin, 1985; 803.
15. Patera AT. A spectral element method for fluid dynamics: laminar flow in a channel expansion. Journal of
Computational Physics 1984; 54: 468–488.
16. Manwell JF, McGowan JG, Rogers A. Wind Energy Explained: Theory, Design and Application. Wiley: Chichester,
2002.
17. Burton T, Sharpe D, Jenkins N, Bossanyi E. Wind Energy Handbook. Wiley: Chichester, 2001.
18. IEC. Wind turbine generator systems—part 13: measurement of mechanical loads. Report IEC/TS 61400-13,
International Electrotechnical Commission (IEC), 2001.
19. Glauert H. Airplane propellers. Aerodynamic Theory 1935; 4: 169–360.
20. Glauert H. A general theory of the autogyro. British ARC 1926; 1111: 558–593.
21. Pitt DM, Peters DA. Theoretical prediction of dynamic-inflow derivatives. Vertica 1981; 5(1): 21–34.

Wind Energ. (2013) © 2013 John Wiley & Sons, Ltd.


DOI: 10.1002/we
A. D. Otero, F. L. Ponta and L. I. Lago Vibrational analysis of the flexo-torsional modes of the NREL 5-MW RWT

22. Leishman JG. Principles of Helicopter Aerodynamics. Cambridge University Press: Cambridge, 2006.
23. Bak C, Aagaard Madsen H, Johansen J. Influence from blade–tower interaction on fatigue loads and dynamic (poster),
Wind Energy for the New Millennium, Proceedings of the 2001 European Wind Energy Conference and Exhibition
(EWEC’01), Copenhagen, Denmark, 2001; 2–6.
24. Powles SRJ. The effects of tower shadow on the dynamics of a horizontal-axis wind turbine. Wind Engineering 1983;
7: 26–42.
25. Viterna LA, Janetzke DC. Theoretical and experimental power from large horizontal-axis wind turbines. Technical
Report, Lewis Research Center, National Aeronautics and Space Administration, Cleveland, OH, 1982.
26. Kooijman HJT, Lindenburg C, Winkelaar D, van der Hooft EL. Dowec 6 MW pre-design. Technical Report,
ECN-CX-01-135, Energy Research Center of the Netherlands, Petten, 2003.
27. Lindenburg C. Aeroelastic modelling of the LMH64-5 blade, ECN DOWEC-02-KL-083/0, Petten, December 2002.
28. UpWind project. [Online]. http://www.upwind.eu/. (Accessed May 2012).
29. IEA Wind subtask 2: research for deeper waters. [Online]. http://www.ieawind.org/AnnexXXIII/Subtask2.html.
(Accessed December 2011).
30. Jonkman J, Butterfield S, Passon P, Larsen T, Camp T, Nichols J, Azcona J, Martinez A. Offshore code compar-
ison collaboration within IEA Wind Annex XXIII: phase II results regarding monopile foundation modeling, 2007
European Offshore Wind Conference & Exhibition, Berlin, Germany, NREL-CP-500-42471, 4–6 December 2007.
31. Passon P, Kühn M, Butterfield S, Jonkman J, Camp T, Larsen T. OC3—benchmark exercise of aero-elastic offshore
wind turbine codes. Journal of Physics: Conference Series 2007; 75: 012–071. IOP Publishing.
32. TPI Composites Inc. Parametric study for large wind turbine blades. Report SAND2002-2519, Sandia National
Laboratories, 2002.
33. Jonkman JM, Buhl ML Jr. Fast user’s guide. Technical Report NREL/EL-500-38230, Golden, CO, 2005. [Online].
http://wind.nrel.gov/designcodes/simulators/fast. (Accessed February 2013).
34. Jonkman J, Laino DJ. NWTC design codes ADAMS2AD. [Online]. http://wind.nrel.gov/designcodes/simulators/
adams2ad/. (Accessed May 2012).
35. Griffin DA. Evaluation of design concepts for adaptive wind turbine blades. Report SAND2002-2424, Sandia National
Laboratories, 2002.
36. Locke J, Contreras Hidalgo I. The implementation of braided composite materials in the design of a bend–twist coupled
blade. Report SAND2002-2425, Sandia National Laboratories, 2002.

Wind Energ. (2013) © 2013 John Wiley & Sons, Ltd.


DOI: 10.1002/we

You might also like