You are on page 1of 13

Comput Mech (2012) 50:821833

DOI 10.1007/s00466-012-0772-0

ORIGINAL PAPER

Fluidstructure interaction modeling of wind turbines: simulating


the full machine
Ming-Chen Hsu Yuri Bazilevs

Received: 6 July 2012 / Accepted: 29 July 2012 / Published online: 22 August 2012
Springer-Verlag 2012

Abstract In this paper we present our aerodynamics and


fluidstructure interaction (FSI) computational techniques
that enable dynamic, fully coupled, 3D FSI simulation of
wind turbines at full scale, and in the presence of the
nacelle and tower (i.e., simulation of the full machine).
For the interaction of wind and flexible blades we employ
a nonmatching interface discretization approach, where the
aerodynamics is computed using a low-order finite-elementbased ALE-VMS technique, while the rotor blades are modeled as thin composite shells discretized using NURBS-based
isogeometric analysis (IGA). We find that coupling FEM
and IGA in this manner gives a good combination of efficiency, accuracy, and flexibility of the computational procedures for wind turbine FSI. The interaction between the
rotor and tower is handled using a non-overlapping sliding-interface approach, where both moving- and stationarydomain formulations of aerodynamics are employed. At the
fluidstructure and sliding interfaces, the kinematic and traction continuity is enforced weakly, which is a key ingredient of the proposed numerical methodology. We present
several simulations of a three-blade 5 MW wind turbine,
with and without the tower. We find that, in the case of no
tower, the presence of the sliding interface has no effect on
the prediction of aerodynamic loads on the rotor. From this
we conclude that weak enforcement of the kinematics gives
just as accurate results as the strong enforcement, and thus
enables the simulation of rotortower interaction (as well as
other applications involving mechanical components in relative motion). We also find that the blade passing the tower
M.-C. Hsu Y. Bazilevs (B)
Department of Structural Engineering, University of California,
San Diego, 9500 Gilman Drive, Mail Code 0085, La Jolla,
CA 92093, USA
e-mail: jbazilevs@ucsd.edu

produces a 1012 % drop (per blade) in the aerodynamic


torque. We feel this finding may be important when it comes
to the fatigue-life analysis and prediction for wind turbine
blades.
Keywords NREL 5 MW offshore Wind turbine aerodynamics Fluidstructure interaction ALE-VMS method
Rotortower interaction Full machine Sliding-interface
formulation

1 Introduction
In the last several years there has been a significant increase
in the research centered around aerodynamic and structural
modeling of wind turbines. While most wind turbine aerodynamics and aeroelasticity simulations for engineering design
are performed using low-fidelity methods (see, e.g., [1,2]),
there has been a growing body of research and development
to raise the fidelity of wind turbine simulations. Standalone
3D computational fluid dynamics simulations of wind turbines may be found in [316], while standalone structural
analyses of rotor blades of complex geometry and material composition under assumed or computed off-line wind
load conditions are reported in [1723]. In simulating wind
turbine aerodynamics, only a handful of researchers (see,
e.g., [7,8,13,24]) considered full wind turbine simulations,
where the wind turbine rotor, nacelle, and tower are all modeled. This is due to the additional computational challenges
associated with the simulation of objects in relative
motion.
Recently, a fully coupled fluidstructure interaction (FSI)
simulation methodology for wind turbine rotors was developed in [25]. The simulations therein showed the importance
of FSI modeling for full-scale wind turbine rotors. However,

123

822

the FSI modeling in [25] only considered the interaction


between the airflow and wind turbine rotor, without taking
into account the tower and nacelle. In this paper we present our computational methodology to include the tower and
nacelle into the FSI modeling, which allows us to simulate
the full machine.
To enable the simulation of rotortower interaction, we
adopt a sliding-interface formulation, which was first proposed in [26] and successfully tested on the NREL Phase VI
wind turbine in [24]. We note that in application of the FEM
to flows with moving mechanical components, the ShearSlip Mesh Update Method [2729] and its more general versions [30,31] may also be employed to handle objects in
relative motion. To handle the FSI we use a nonmatching
interface discretization developed in [32], where we couple
low-order FEM for aerodynamics and Isogeometric Analysis
(IGA) [33,34] for structural mechanics. Using nonmatching
discretizations at the fluidstructure interface can serve purposes beyond an implementation convenience typically associated with weakly-coupled FSI methods. What we do here
is an example of taking advantage of the flexibility associated with using nonmatching discretizations, combining the
most appropriate discretization for each part of an FSI problem. Exploiting this kind of flexibility, in conjunction with
FSI homogenization methods [3537], enabled computation
of some of the most challenging parachute FSI problems
[3739], including parachute clusters [38,39], where the
contact model [39,40] needed to deal with the interaction between the parachutes took advantage of this flexibility. Turbulence modeling in the aerodynamics formulation makes
use of the recently proposed ALE-VMS technique [41,42] in
combination with weakly enforced essential boundary conditions [43]. The latter greatly improve the accuracy of the
ALE-VMS formulation in the presence of thin boundary layers near the rotor surface [44]. The KirchhoffLove shell formulation [45] in conjunction with the bending strip method
[21] is used to model blade structures made of composite
materials.
This paper is outlined as follows. In Sect. 2, we first state
the continuum formulation of the FSI problem suitable for
a nonmatching discretization of the fluidstructure interface.
The fluid and structural mechanics formulations at the spacediscrete level are shown next, followed by the presentation
of the sliding-interface formulation. We conclude the section
with a brief discussion of the time integration and FSI coupling techniques. In Sect. 3, we present the numerical results
for the NREL 5 MW offshore wind turbine [2,9]. We first
verify that the aerodynamic loads are insensitive to the presence of the sliding interface. We then compute several cases
to assess the importance of including FSI and rotortower
interaction in the modeling. It is shown that both produce
an appreciable effect. In Sect. 4, we draw conclusions and
outline future developments.

123

Comput Mech (2012) 50:821833

2 Continuum and discrete modeling of FSI


In this section, we present the computational framework
for FSI analysis of a full wind turbine. For the FSI formulation we use the developments in [32]. The coupled FSI
formulation is able to handle nonmatching fluidstructure
interface discretizations without introducing the interface
degrees-of-freedom. This is done for better efficiency of the
computational procedures. To take into account the interaction between the rotor and tower, a non-overlapping sliding-interface method [26,24] that handles subdomains in
relative motion is also presented. The sliding-interface discretization also avoids the use of interface degrees-offreedom.

2.1 FSI formulation at the continuum level suitable


for nonmatching interface discretizations
Let (t )1 Rd , d = 2, 3, represent the time-dependent
domain of the fluid mechanics (or aerodynamics) problem
with boundary (t )1 , let (t )2 Rd represent the timedependent domain of the structural mechanics problem with
boundary (t )2 , and let (t )I Rd represent the interface
between the fluid and structural domains. Let u1 and p denote
the fluid velocity and pressure, respectively, and let u2 denote
the velocity of the structure. Let Su , S p , and Sd be the trial
function spaces for the fluid velocity and pressure, and structural velocity, respectively, and Vu , V p , and Vd be the corresponding test function spaces. The variational formulation
for the coupled problem may be stated as: find u1 Su , p
S p , and u2 Sd , such that w1 Vu , q V p , and w2
Vd ,

B1 ({w1 , q}, {u1 , p}; u)F


1 ({w1 , q})+B2 (w2 , u2 )F2 (w2 )


(w1 w2 ) 1 (u1 , p) n1 d
(t )I


1 (w1 , q) n1 (u1 u2 ) d

(t )I


+

(w1 w2 ) (u1 u2 ) d = 0.

(1)

(t )I

In the above, the subscripts 1 and 2 denote the fluid and


structural mechanics quantities, and is a penalty parameter, which we leave unspecified for the moment. Furthermore,
B1 , F1 , B2 , and F2 are the semilinear forms and linear functionals corresponding to the fluid and structural mechanics
problems, and are given by

Comput Mech (2012) 50:821833

823



B1 {w, q}, {u, p}; u

 



u 

=
w 1
+
u

u
d
t x
(t )1

formulation given by Eq. (1) reduces to a form suitable for


matching fluidstructure interface meshes. Although somewhat limiting, matching interface discretizations were successfully applied to cardiovascular FSI in [4957].


(w) : 1 (u, p) d +

+
(t )1

w 1 f1 d +
(t )1

w 2

du
d +
dt

(w) : 2 (u) d,

(t )2

(4)


w 2 f2 d +

(t )2

(3)

(t )1h


F2 (w) =

w h1 d,

2.2 ALE-VMS formulation of fluid mechanics with weak


boundary conditions

F1 ({w, q}) =

(t )2

(2)

(t )1

B2 (w, u) =

u d,
q

w h2 d,

(5)

(t )2h

where s are the densities, fs are the prescribed body forces,


hs are the prescribed surface tractions, u is the velocity of
the fluid domain (t )1 , (t )1h and (t )2h are the boundaries

where the surface tractions are specified,  denotes the fact
x
that the time derivative in the fluid mechanics equations is
taken with respect to the fixed referential domain spatial coord()
denotes the total time derivative (i.e., time
dinates x , and
dt
derivative taken with respect to the fixed coordinates of the
material configuration). The fluid mechanics Cauchy stress
tensor 1 is defined as
1 (u, p) = pI + 2 (u) ,

(6)

where I is the identity tensor, is the dynamic viscosity, and


(u) is the strain-rate tensor given by

1
u + uT .
(7)
(u) =
2
In Eq. (1),

We assume that the time-dependent fluid domain (t )1 is


 into Nel spatial subdomain elements denoted by
divided
e
t 1 . The fluidstructure interface (t )I is decomposed into
Neb fluid domain surface elements denoted by tb , and (t )
I
is defined as the inflow part of (t )I . The discrete trial function spaces Suh for the velocity and S hp for the pressure, as
well as the corresponding test function spaces Vuh and V hp are
assumed to be of equal order, and, in this work, are comprised
of FEM functions. The discrete velocities and pressures and
the corresponding test functions are now superscripted with
h to denote their dependence on the mesh size.
The ALE-VMS formulation of the fluid mechanics subproblem, including the weakly enforced boundary conditions, is obtained by setting w2 = 0, B1 = B1VMS , F1 =
F1VMS , and = B in Eq. (1), and is given by: find u1h Suh
and p h S hp , such that w1h Vuh and q h Vhp ,
B1VMS

At this point we do not detail the structural Cauchy stress


2 as we wish to keep the structural mechanics formulation
flexible.

Remark 2 If the fluid and structural velocities and the corresponding test functions are explicitly assumed to be continuous (i.e., u1 = u2 and w1 = w2 ) at the interface, the FSI

Neb

b=1 b
t (t )I

(8)

Remark 1 The variational formulation given by Eq. (1) may


be interpreted as Nitsches method applied to the FSI problem
(see, e.g., [46,47]), or as a continuous version of the discontinuous Galerkin (DG) method (see, e.g., [48]) applied at the
fluidstructure interface. It is assumed in Eq. (1) that the trial
and test function spaces of the fluid and structural subproblems are independent of each other. This approach provides
a framework that is capable of handling nonmatching fluid
and structural interface discretizations.

Neb

b=1 b
t (t )I

1 (w, q) n = 2 (w)n + qn.


w1h , q h , u1h , p h ; u h F1VMS w1h , q h


w1h 1 u1h , p h n1 d




2 w1h n1 + q h n1 u1h u h d

Neb

b=1 b
t (t )
I

Neb

b=1 b
t (t )I

w1h 1





u1h u h n1 u1h u h d



B w1h u1h u h d = 0,

(9)

where



wh , q h , uh , p h ; u h




= B1 wh , q h , uh , p h ; u h

B1VMS

Nel 

e=1

(et )1
Nel 







qh
wh +
uh u h
rM uh , p h d
1



1 C wh rC uh , p h d

e=1

(et )1

123

824

Comput Mech (2012) 50:821833

Nel 



 
M wh rM uh , p h uh d

e=1

(et )1


Nel

e=1

(et )1





wh 
: M rM uh , p h M rM uh , p h d,
1

(10)
and F1VMS

wh , q h

= F1


wh , q h .

(11)

In the above equations, u h is the mesh velocity, and rM and rC


are the residuals of the momentum and continuity equations,
respectively, given by
 h





u 
h
h
h
h
h
h
+ u u u f
rM u , p = 1
t x


1 uh , p h ,

(12)
and


r C uh , p h = uh .

(ts )2

(13)

The definitions of M and C , the stabilization parameters,


may be found in [5866,50,67]).
The ALE-VMS method is a moving-domain extension of
the residual-based variational multiscale (RBVMS) formulation proposed in [66] in the context of nonmoving domain
problems. Weakly enforced essential boundary conditions
were introduced in [43], which improved the performance of
the ALE-VMS formulation in the presence of underresolved
boundary layers [6870]. The definition of the boundary stabilization (or penalty) parameter B may be found in [44].
In the case of FSI, the essential boundary conditions for the
fluid mechanics problem at the fluidstructure interface come
from the structural velocity. In this case, the mesh velocity
u h on (t )I may be computed as
u h = 1h u2h ,

(14)

where 1h is a projection or interpolation operator onto the


space spanned by the basis functions of the fluid mechanics problem restricted to the fluidstructure interface. In this
work we use an L 2 -projection. Equation (14) gives the coupling to the structural mechanics problem, which is discussed
in the next section.
2.3 KirchhoffLove shell formulation of structural
mechanics and the bending strip method
The discrete variational formulation of the structural mechanics problem is obtained by setting {w1 , q} = {0, 0} in Eq. (1)
and assuming that the structure is governed by the isogeometric KirchhoffLove composite shell formulation with the aid
of the bending strip method [45,21,25]. The shell midsurface
is described using a piecewise smooth (C 1 - or higher-order

123

continuous) geometrical mapping. Regions where the mapping reduces to the C 0 level are also allowed. This includes
situations where the shell surface geometry is described using
several NURBS patches that are joined with C 0 -continuity.
For this reason, we define (2s )0 and (2s )t to be the structure midsurface, which is composed of the structural patches
or subdomains, in the reference and deformed configuration,
respectively. Furthermore, (2b )0 denotes the bending strip
domain, which is a union of the bending strip patch subdomains, in the reference configuration.
Let Sdh and Vdh denote the discrete trial and test function
spaces for the structural mechanics problem. The rotationfree KirchhoffLove shell formulation may be stated as: find
the velocity of the shell midsurface u2h Sdh , such that for
all test functions w2h Vdh ,



du2h
h
f d
w2 2 h th
dt


+
(0s )2

+
(0s )2



h A h + B h d



h
h
h

d +
B + D


w2h 2h t1h d = 0.

h Db h d

(0b )2

(15)

(t )I

In the above, h th is the shell thickness, h and h are the


tensors of membrane strains and curvature changes in Voigt notation written with respect to the local Cartesian basis
oriented on the first covariant basis vector of the midsurface, h and h are their variations, A, B and D are the
membrane, coupling and bending stiffnesses, respectively,
pre-integrated through the thickness, Db is the bending stiffness of the bending strips, 2h is an L 2 -projection operator
onto the space spanned by the basis functions of the structural
mechanics problem restricted to the fluidstructure interface,
and t1h is the discrete fluid traction given by


1h n1 B u2h u1h .
(16)
t1h =
Although NURBS are employed in this work to discretize the
structural mechanics equations, T-splines [71,72] or subdivision surfaces [7375] are also well suited for the proposed
structural modeling approach. For alternative rotation-free
shell formulations the reader is referred to [7379].
2.4 Sliding-interface formulation for objects in relative
motion
In order to simulate the full wind turbine configuration and
investigate the rotortower interaction, we consider an

Comput Mech (2012) 50:821833

825

Neb

b=1 b
t (t )SI

Neb

b=1 b
t (t )SI

Neb

Neb

b=1 b
t (t )SI

Fig. 1 Setup for the simulation of a full machine. The interior moving subdomain, which encloses the wind turbine rotor, and the exterior
stationary subdomain, which houses the nacelle and tower, are shown

approach that makes use of a moving subdomain, which


encloses the entire wind turbine rotor, and a stationary subdomain that contains the rest of the wind turbine (see Fig. 1). The
two domains are in relative motion and share a sliding cylindrical interface. The meshes on each side of the interface are
nonmatching because of the relative motion (see Fig. 2). As
a result, a numerical procedure is needed to impose the continuity of the kinematics and tractions at the stationary and
rotating subdomain interface despite the fact that the interface discretizations are incompatible. Such a procedure was
developed in [26] in the context of IGA for computing flows
about rotating components. The advantage of IGA for rotating-component flows is that the cylindrical sliding interfaces
are represented exactly and no geometry errors are incurred.
In the case of standard FEM employed here, the geometric
compatibility is only approximate. The sliding-interface coupling was successfully tested on the NREL Phase VI wind
turbine in [24] and is presented in what follows.
Let the subscripts 1S and 1M denote the quantities pertaining to the fluid mechanics problem on the stationary and moving subdomains, respectively. The subdomain that encloses
the rotor rotates with it, and the interior of the rotating subdomain is allowed to deflect to accommodate the motion of
the blades. However, the motion of the outer boundary of the
rotor subdomain is restricted to a rigid rotation to maintain
geometric compatibility with the stationary subdomain. To
enforce the compatibility of the flow kinematics and tractions at the sliding interface, we add the following terms to
the ALE-VMS formulation given by Eq. (9):

Neb

b=1 b
t (t )SI

1
h
h
1S n1S 1M n1M ) d
w1S
(
w1M
2

h
w1S

b=1 b
t (t )SI



1
h
h
1S n1S
1M n1M ) u1S
u1M
d
(
2

h
w1M




h
h
h
h
n1S
u1S
d
u1S
u 1S
u1M




h
h
h
h
n1M
u1M
d
u1M
u 1M
u1S



CIB  h
h
h
h
u1M
w1S w1M
u1S
d = 0,
hn

(17)

is given by Eq. (8), (t )SI is the sliding interface,


where
and {A} denotes the negative part of A, that is, {A} = A if
A < 0 and {A} = 0 if A 0. Just as in the FSI case, the sliding-interface formulation may be see as a DG method, where
the continuity of the basis function is enforced everywhere
in the interior of the two subdomains, but not at the sliding
interface between them. The structure of the terms on the
sliding interface is similar to that of the weak enforcement
of essential boundary conditions (see Eq. (9)). The significance of each term is explained in detail in [26]. Note that,
h = 0, because the subdomain
in the current application, u 1S
1S is stationary. However, the formulation is able to handle
situations where both subdomains are in motion.
Remark 3 Nonmatching interface discretizations in the FSI
and sliding-interface problems necessitate the use of interpolation or projection of kinematic and traction data
between the nonmatching surface meshes (see, e.g., [80,81,
32,82], where [82] is more comprehensive than [81]). A computational procedure, which can simultaneously handle the
data transfer for IGA and FEM discretizations, was proposed
in [32]. The procedure also includes a robust approach in
identifying closest points for arbitrary shaped surfaces.
While such interface projections are rather straightforward
for weakly-coupled FSI algorithms, they require special techniques [65,40,82] for strongly-coupled, direct and quasidirect methods [83,84,65,40,82] that are monolithic-like
(i.e. become monolithic for matching discretizations).
2.5 Time integration of the FSI equations and coupling
The ALE formulation for fluid mechanics and the Lagrangian formulation for structural mechanics give a natural setting for a finite-difference time integration of this coupled
FSI system. In this work, we employ the Generalized- technique [85,86,50], which is a fully-implicit second-order
accurate method with control over the dissipation of high-frequency modes. At each time step the combined fluid, structure and mesh motion discrete residuals are converged to zero
by means of a block-iterative FSI coupling approach [83,84,
65]. The block-iterative approach for nonconforming fluid

123

826

Comput Mech (2012) 50:821833

Fig. 2 Nonmatching meshes at the sliding interface between the stationary and moving subdomains

and structural meshes consists of the following sequence of


linear solves. We first compute the fluid mechanics solution
increment holding the structural and mesh solutions fixed. We
then update the fluid solution, extract the fluid traction and
project it to the structure, and obtain the structural mechanics solution increment. Finally, we update the structural solution, project the structural kinematics data to the fluid domain
boundary, and update the mesh solution increment using the
linear elasticity operator. This three-step iteration sequence
is repeated until convergence to an appropriately coupled
discrete solution is achieved. We note that only the mesh
motion part associated with blade deflections is computed
using the mesh moving method that we typically use, which
is an elastic mesh moving technique with Jacobian-based
stiffening [8789]. The mesh motion part associated with
blade rotations is computed exactly (see [25] for details).
The block-iterative approach to FSI coupling is stable and
efficient for wind turbine simulation.

3 Computational results
In this section, we present our computations of the NREL
5 MW offshore baseline wind turbine [2,9]. The NREL 5 MW
wind turbine is a conventional three-blade upwind turbine
proposed in [2] to support concept studies aimed at assessing
offshore wind technology [9096]. The rotor-only configuration of this wind turbine was simulated earlier in
[9,25,10,11,97,32,41,44] using both NURBS-based IGA
and standard FEM. The detailed geometry description and
construction for the blade surface was documented in [9].
The 63 m blade is composed of a series of DU airfoils and
the NACA64 profile. In this work, the proposed framework

123

is applied to the simulation of the full wind turbine configuration, including the rotor (blades and hub), nacelle, and
tower. The land-based tower is assumed to be rigid, and has
a base diameter of 6 m and a top diameter of 3.87 m. The
tower height is 87.6 m and the hub height is 90 m [2].
The cases considered in this work are show in Fig. 3. Case I
is a rotor-only setup, where the rotor is enclosed in a cylindrical domain, and the rotation is applied to the entire computational domain to simulate the spinning rotor. Case II is also a
rotor-only setup, however, the rotor is housed in a cylindrical
rotating subdomain that is enclosed by a stationary exterior
flow subdomain. The sliding-interface formulation is applied
at the interface between the subdomains to weakly impose
the continuity of the kinematics and tractions. We compare
the simulation results from Cases I and II to study the effect
of the sliding-interface formulation. Case III has a similar
setup as Case II, but this time the stationary subdomain contains the nacelle and tower, and thus enables us to study the
rotortower interaction effects.
All simulations are performed at the rated wind speed of
11.4 m/s and a fixed rotor speed of 12.1 rpm. This setup corresponds to one of the cases reported in [2]. The wind speed is
prescribed at the inflow boundary, the traction vector is set to
zero at the outflow boundary, and the slip condition is set on
the top, bottom, and lateral boundaries. The air density and
viscosity are 1.2 kg/m3 and 2.0105 kg/(m s), respectively.
The aerodynamics volume mesh statistics for all three
cases are summarized in Table 1. Case III has more elements
and nodes compared with Cases I and II due to the presence
of the nacelle and tower. A 2D cut of the mesh at x = 0 for
Case III is shown in Fig. 4 to show the mesh quality used
in our computations. The mesh is refined in the inner region
for better flow resolution. Figure 5 shows a 2D cut at 75 %

Comput Mech (2012) 50:821833

827

Fig. 3 Cases considered. (a) Case I: rotor only. (b) Case II: rotor only, including a sliding interface. (c) Case III: full machine with a sliding
interface

Table 1 Mesh statistics for Cases I, II and III


Number of nodes

Number of elements

Case I

1,319,427

4,218,459

Case II

1,326,129

4,205,519

Case III

1,440,425

4,828,692

Fig. 5 A 2D cut at 75 % spanwise station to illustrate the boundarylayer mesh. The size of the first element in the wall-normal direction
is 0.02 m, and a growth ratio of 1.2 is used to generate 15 layers of
prismatic elements

3.1 Rigid-blade simulations

Fig. 4 A 2D cut at x = 0 to show the mesh quality used in Case III.


The mesh is refined in the inner region for better flow resolution near
the wind turbine

spanwise station to illustrate the boundary-layer mesh used in


our computations. Near the blade surface, the size of the first
element in the wall-normal direction is 0.02 m, and 15 layers
of prismatic elements were generated with a growth ratio of
1.2. The Reynolds number based on the chord length and
relative speed at this location is O(107 ). The same boundary-layer mesh is used in all three cases. The time-step size
is 2.5 104 s for all cases. Both rigid- and flexible-blade
computations are performed for each case.

We first perform the simulations assuming the rotor blades


are rigid. The time history of the aerodynamic torque is plotted in Fig. 6 for all three cases. The comparison between Case
I and II shows that the presence of the sliding interface has
no effect on the prediction of aerodynamic loads on the rotor.
From this we conclude that weak enforcement of the kinematics at the sliding interface gives just as accurate results as
the strong enforcement, and thus enables the simulation of
rotortower interaction (as well as other applications involving mechanical components in relative motion).
Figure 6 also shows the time history of the aerodynamic
torque for Case III, where the nacelle and tower are included
and the full machine is considered in the simulation. The
results clearly show several dips in the torque history, each
dip corresponding to one of the blades passing the tower.
To better examine the tower effect, we plot the aerodynamic
torque for each individual blade as a function of the azimuthal
angle (see Fig. 7). For each blade, the azimuthal angle of 180

123

828

Comput Mech (2012) 50:821833


Azimuthal angle of blade #1 (degree clockwise)

Azimuthal angle of blade #1 (degree clockwise)


0

145.2 290.4 75.6 220.8

6.0

151.2 296.4 81.6 226.8 12.0

220.8
5000

Case I:
Rigid Blade
Rotor Only

4500

Aerodynamic torque (kN m)

Aerodynamic torque (kN m)

5000

4000
3500
3000
2500
2000

Case II:
Rigid Blade
Rotor Only
(Sliding Interface)

Case III:
Rigid Blade
Full Machine
(Sliding Interface)

10

12

14

16

18

151.2

296.4

81.6

Rigid Blade
Full Machine
(Sliding Interface)

4500

226.8

12.0

157.2

Case I:
Flexible Blade
Rotor Only

4000
3500

Case II:
Flexible Blade
Rotor Only
(Sliding Interface)

3000

Case III:
Flexible Blade
Full Machine
(Sliding Interface)

2500
2000

20

6.0

Time (s)

10

12

14

16

18

20

22

Time (s)

Fig. 6 The aerodynamic torque for the rigid-blade computations. For


the full-machine simulation, the torque drops when the blade passes in
front of the tower

Fig. 8 The aerodynamic torque for flexible-blade FSI computations


Azimuthal angle of blade #1 (degree clockwise)
220.8
-6

1300
1250
1200
1150
1100
1050

6.0

151.2

296.4

81.6

Blade #3

1350

Blade tip z-displacement (m)

Aerodynamic torque (kN m)

1400

60

120

180

240

300

157.2

-4
-3
-2
-1

360

Azimuthal angle (degree clockwise)

Fig. 7 The single-blade aerodynamic torque as a function of the azimuthal angle for a rigid-blade full-machine simulation. The 180 angle
corresponds to the instant when the blade passes in front the tower. At
that instant the predicted drop in the torque is 1012 % of its computed
value when the blade is away from the tower

represents the instant when the blade passes the tower. At


that instant the predicted drop in the torque is 1012 % of its
computed value when the blade is away from the tower. This
finding may be important when it comes to the fatigue-life
analysis and prediction for wind turbine blades.
3.2 Flexible-blade simulations
The rotor blades are discretized using NURBS-based IGA.
A symmetric fiberglass/epoxy composite with [45/0/902 /
03 ]s lay-up was proposed in [25] for the blade material. The
details of the NURBS structure model, including the homogenization of the material and the laminate thickness distributions, can be found in [25,98]. There are 4,897 control points
in the quadratic NURBS mesh.

123

12 . 0

Blade #2

Blade #1

-5

226.8

10

12

14

16

18

20

22

Time (s)

Fig. 9 The blade tip z-displacement for rotor-only (dashed, dasheddotted, and dashed-dotted-dotted lines) and full-machine (solid lines)
simulations. The arrows near the bottom of the figure show the instants
where the blade passes in front of the tower

We perform FSI (flexible-blade) simulations for all three


cases. The time history of the aerodynamic torques are shown
in Fig. 8. The comparison between Cases I and II again shows
that the presence of the sliding interface has no effect on
the prediction of aerodynamic loads on the rotor. We also
note the transient dynamic effect as well as high frequency
oscillations in the FSI curves compared with the rigid-blade
simulation result. The tower effect is still appreciable.
Figure 9 shows the tip z-displacement history of the individual blades. The solid lines are the results from Case III
(full machine), while the dashed, dashed-doted, and dasheddoted-doted lines are from Case II (rotor only). From the
comparison we note that the rotortower interaction doesnt
seem to produce a pronounced response in the blade tip displacement.

Comput Mech (2012) 50:821833

829
Azimuthal angle of blade #1 (degree clockwise)

Azimuthal angle of blade #1 (degree clockwise)


6.0

151.2

296.4

Blade #2

1500

81.6

226.8

12.0

Blade #3

1250

1000

Blade #1
750

500

220.8
1750

157.2

Aerodynamic torque (kN m)

Aerodynamic torque (kN m)

220.8
1750

10

12

14

16

18

20

151.2

296.4

226.8

12.0

157.2

1500

1250

1000

Full Machine
Blade #1

750

22

10

12

14

16

18

20

22

Time (s)

Time (s)

Azimuthal angle of blade #1 (degree clockwise)

Fig. 10 The aerodynamic torque of individual blades for flexible-blade


full-machine simulation
Aerodynamic torque (kN m)

220.8
1750

6.0

151.2

296.4

81.6

226.8

12.0

157.2

Rotor Only
Blade #2

1500

1250

1000

Full Machine
Blade #2

750

500

10

12

14

16

18

20

22

Time (s)
Azimuthal angle of blade #1 (degree clockwise)
220.8
1750

Aerodynamic torque (kN m)

Figure 10 shows the aerodynamic torques of individual


blades for Case III. Note that the torque histories are clearly
different between three blades. From Figs. 9 and 10 we also
note that the individual tip displacement and torque do not
settle at the steady values. Once the initial transient decays,
the tip displacement varies between 3 and 4 m during one
revolution. The dynamic effect is also present in individual torque curves shown in Fig. 10. After the initial transient
decays, the torque varies between 1,180 and 1,330 kN m during one revolution. Figure 11 compares the torque history of
individual blades between Cases II and III to show the effect
of rotortower interaction in FSI computations. The drop in
aerodynamic torque is clearly seen for all three blades.
Figure 12 shows the relative air speed and streamlines at
50 % spanwise station, rotated to the reference configuration,
and superposed on the moving blade #2. The blade mostly
displaces in the flapwise direction, however, the edgewise
displacement may also be observed. The weak enforcement
of boundary conditions results in the flow slipping on the
solid surface. Figure 13 shows the flow visualization of the
full wind turbine configuration. The tip vortex decays very
slowly as it is convected downstream. Note that no visible
discontinuities are seen in the flow field at the sliding interface, which indicates that the method correctly handles the
kinematic compatibility at this location.
The time-averaged aerodynamic torque values for all cases
are summarized in Table 2. From the FSI computation of Case
III, and without considering any power loss, the energy generated by this wind turbine design is about 4.73 MW. According to the Betz law (see, e.g. [99]), the maximum power that
this wind turbine design is able to extract for the wind and
rotor speeds considered is 6.57 MW, leading to the aerodynamic efficiency of 72 %, which is quite good for modern
wind turbine designs.

81.6

Rotor Only
Blade #1

500
8

6.0

6.0

151.2

296.4

81.6

226.8

12.0

157.2

18

20

22

Rotor Only
Blade #3

1500

1250

1000

Full Machine
Blade #3

750

500

10

12

14

16

Time (s)

Fig. 11 The aerodynamic torque of individual blades for rotor-only


and full-machine simulations (flexible blade).The drop in the aerodynamic torque is clearly visible

4 Conclusions
We presented our aerodynamics and FSI computational techniques that enable dynamic, fully coupled, 3D FSI simulation

123

830

Comput Mech (2012) 50:821833

Fig. 12 Relative air speed and streamlines at 50 % spanwise station, rotated to the reference configuration, and superposed on the moving blade
#2 at three different time instants: (a) t = 9.0 s, (b) t = 10 s, and (c) t = 12.5 s

Fig. 13 Air speed contours at a planar cut (left) and isosurfaces of air speed (right) at an instant for the flexible-blade full-machine simulation

Table 2 Time-averaged (over one revolution) aerodynamic torques


Rigid-blade
aerodynamic
torque (kN m)

Flexible-blade
aerodynamic
torque (kN m)

Case I

3,735

3,749

Case II

3,733

3,751

Case III

3,716

3,734

of wind turbines at full scale, and in the presence of the nacelle and tower (i.e., simulation of the full machine). The
structural modeling of wind turbine blades makes use of the
isogeometric rotation-free shell formulation, which is coupled to the low-order finite-element-based ALE-VMS technique for aerodynamics. This presents a good combination of
accuracy and efficiency of the wind turbine FSI simulations.

123

The bladetower interaction is successfully handled using a


sliding-interface technique with weak enforcement of flow
kinematic and traction compatibility.
The computational results indicate that the blade passing
the tower produces an appreciable drop in the aerodynamic
torque. This creates additional cyclic loading on the blades
that may be important when it comes to their fatigue-life analysis. Furthermore, the FSI modeling employed in this work
captures the variability in the deformation of the individual
blades as well as that of the aerodynamic torque during one
revolution. Because at any instant in time the blade deformation is nonsymmetric, this creates eccentric loads on the
hub, which may be important to predict for the purposes of
hub and nacelle design.
Acknowledgments The support of the NSF CAREER Award
No. 1055091 is gratefully acknowledged. We thank the Texas Advanced
Computing Center (TACC) at the University of Texas at Austin and the

Comput Mech (2012) 50:821833


San Diego Supercomputing Center (SDSC) at the University of California, San Diego for providing HPC resources that have contributed to
the research results reported in this paper.

References
1. Jonkman JM, Buhl ML Jr (2005) FAST users guide. Technical
report NREL/EL-500-38230. National Renewable Energy Laboratory, Golden
2. Jonkman J, Butterfield S, Musial W, Scott G (2009) Definition
of a 5-MW reference wind turbine for offshore system development. Technical Report NREL/TP-500-38060. National Renewable Energy Laboratory, Golden,CO
3. Srensen NN, Michelsen JA, Schreck S (2002) NavierStokes predictions of the NREL phase VI rotor in the NASA Ames 80 ft
120 ft wind tunnel. Wind Energy 5:151169
4. Duque EPN, Burklund MD, Johnson W (2003) NavierStokes and
comprehensive analysis performance predictions of the NREL
Phase VI experiment. J Solar Energy Eng 125:457467
5. Le Pape A, Lecanu J (2004) 3D NavierStokes computations of a
stall-regulated wind turbine. Wind Energy 7:309324
6. Chao DD van Dam CP (2007) Computational aerodynamic analysis of a blunt trailing-edge airfoil modification to the NREL Phase
VI rotor. Wind Energy 10:529550
7. Gmez-Iradi S, Steijl R, Barakos GN (2009) Development and validation of a CFD technique for the aerodynamic analysis of HAWT.
J Solar Energy Eng 131:031009-1-13
8. Zahle F, Srensen NN, Johansen J (2009) Wind turbine rotor-tower
interaction using an incompressible overset grid method. Wind
Energy 12:594619
9. Bazilevs Y, Hsu MC, Akkerman I, Wright S, Takizawa K, Henicke
B, Spielman T, Tezduyar TE (2011) 3D simulation of wind turbine
rotors at full scale. Part I: Geometry modeling and aerodynamics.
Int J Num Methods Fluids 65:207235
10. Takizawa K, Henicke B, Tezduyar TE, Hsu MC, Bazilevs
Y (2011) Stabilized spacetime computation of wind-turbine
rotor aerodynamics. Comput Mech 48:333344
11. Takizawa K, Henicke B, Montes D, Tezduyar TE, Hsu M-C,
Bazilevs Y (2011) Numerical-performance studies for the stabilized spacetime computation of wind-turbine rotor aerodynamics.
Comput Mech 48:647657
12. Bechmann A, Srensen NN, Zahle F (2011) CFD simulations of
the MEXICO rotor. Wind Energy 14:677689
13. Li Y, Carrica PM, Paik KJ, Xing T (2012) Dynamic overset CFD
simulations of wind turbine aerodynamics. Renew Energy 37:285
298
14. Srensen NN, Schreck S (2012) Computation of the national
renewable energy laboratory phase-VI rotor in pitch motion during
standstill. Wind Energy. doi:10.1002/we.480
15. Scheurich F, Brown RE (2012) Modelling the aerodynamics of vertical-axis wind turbines in unsteady wind conditions. Wind Energy.
doi:10.1002/we.532
16. Chow R, van Dam CP (2012) Verification of computational simulations of the NREL 5 MW rotor with a focus on inboard flow
separation. Wind Energy. doi:10.1002/we.529.
17. Kong C, Bang J, Sugiyama Y (2005) Structural investigation of
composite wind turbine blade considering various load cases and
fatigue life. Energy 30:21012114
18. Lund E, Stegmann J (2005) On structural optimization of composite shell structures using a discrete constitutive parametrization.
Wind Energy 8:109124
19. Hansen MOL, Srensen JN, Voutsinas S, Srensen N, Madsen Aa
H (2006) State of the art in wind turbine aerodynamics and aeroelasticity. Prog Aerosp Sci 42:285330

831
20. Jensen FM, Falzon BG, Ankersen J, Stang H (2006) Structural testing and numerical simulation of a 34 m composite wind turbine
blade. Compos Struct 76:5261
21. Kiendl J, Bazilevs Y, Hsu M-C, Wchner R, Bletzinger K-U
(2010) The bending strip method for isogeometric analysis of KirchhoffLove shell structures comprised of multiple patches. Comput Methods Appl Mech Eng 199:24032416
22. Bazilevs Y, Hsu M-C, Kiendl J, Benson DJ (2012) A computational procedure for prebending of wind turbine blades. Int J Num
Methods Eng 89:323336
23. Crdenas D, Escrpita AA, Elizalde H, Aguirre JJ, Marzocca P,
Ahuett H, Probst O (2012) Numerical validation of a finite element thin-walled beam model of a composite wind turbine blade.
Wind Energy 15:203223
24. Hsu M-C, Akkerman I, Bazilevs Y (2012) Finite element simulation of wind turbine aerodynamics: validation study using NREL
Phase VI experiment. Wind Energy (In review)
25. Bazilevs Y, Hsu M-C, Kiendl J, Wchner R, Bletzinger K-U
(2011) 3D simulation of wind turbine rotors at full scale. Part II:
Fluidstructure interaction modeling with composite blades. Int J
Num Methods Fluids 65:236253
26. Bazilevs Y, Hughes TJR (2008) NURBS-based isogeometric analysis for the computation of flows about rotating components. Comput Mech 43:143150
27. Tezduyar T, Aliabadi S, Behr M, Johnson A, Kalro V, Litke
M (1996) Flow simulation and high performance computing. Comput Mech 18:397412
28. Behr M, Tezduyar T (1999) The shear-slip mesh update method.
Comput Methods Appl Mech Eng 174:261274
29. Behr M, Tezduyar T (2001) Shear-slip mesh update in 3D computation of complex flow problems with rotating mechanical components. Comput Methods Appl Mech Eng 190:31893200
30. Tezduyar TE (2001) Finite element methods for flow problems
with moving boundaries and interfaces. Arch Comput Methods
Eng 8:83130
31. Tezduyar TE (2007) Finite elements in fluids: special methods and
enhanced solution techniques. Comput Fluids 36:207223
32. Bazilevs Y, Hsu MC, Scott MA (2012) Isogeometric fluidstructure interaction analysis with emphasis on non-matching discretizations, and with application to wind turbines. Comput Methods
Appl Mech Eng. doi:10.1016/j.cma.2012.03.028
33. Hughes TJR, Cottrell JA, Bazilevs Y (2005) Isogeometric analysis: CAD finite elements, NURBS exact geometry, and mesh refinement. Comput Methods Appl Mech Eng 194:41354195
34. Cottrell JA, Hughes TJR, Bazilevs Y (2009) Isogeometric analysis:
toward integration of CAD and FEA. Wiley, Chichester
35. Tezduyar TE, Sathe S, Pausewang J, Schwaab M, Christopher J,
Crabtree J (2008) Interface projection techniques for fluidstructure interaction modeling with moving-mesh methods. Comput
Mech 43:3949
36. Tezduyar TE, Sathe S, Pausewang J, Schwaab M, Christopher J,
Crabtree J (2008) Fluidstructure interaction modeling of ringsail
parachutes. Comput Mech 43:133142
37. Takizawa K, Moorman C, Wright S, Spielman T, Tezduyar
TE (2011) Fluidstructure interaction modeling and performance
analysis of the Orion spacecraft parachutes. Int J Num Methods
Fluids 65:271285
38. Takizawa K, Wright S, Moorman C, Tezduyar TE (2011) Fluid
structure interaction modeling of parachute clusters. Int J Num
Methods Fluids 65:286307
39. Takizawa K, Spielman T, Tezduyar TE (2011) Spacetime FSI
modeling and dynamical analysis of spacecraft parachutes and
parachute clusters. Comput Mech 48:345364
40. Takizawa K , Tezduyar TE (2012) Computational methods for
parachute fluidstructure interactions. Arch Comput Methods Eng
19:125169

123

832
41. Bazilevs Y, Hsu M-C, Takizawa K, Tezduyar TE (2012) ALEVMS
and STVMS methods for computer modeling of wind-turbine
rotor aerodynamics and fluidstructure interaction. Math Models
Methods Appl Sci. doi:10.1142/S0218202512300025
42. Takizawa K, Bazilevs Y, Tezduyar TE (2012) Spacetime and
ALE-VMS techniques for patient-specific cardiovascular fluid
structure interaction modeling. Arch Comput Methods Eng
19:171225
43. Bazilevs Y, Hughes TJR (2007) Weak imposition of Dirichlet
boundary conditions in fluid mechanics. Comput Fluids 36:
1226
44. Hsu M-C, Akkerman I, Bazilevs Y (2012) Wind turbine aerodynamics using ALEVMS: validation and the role of weakly
enforced boundary conditions. Comput Mech. doi:10.1007/
s00466-012-0686-x
45. Kiendl J, Bletzinger K-U, Linhard J, Wchner R (2009) Isogeometric shell analysis with KirchhoffLove elements. Comput
Methods Appl Mech Eng 198:39023914
46. Nitsche J (1971) Uber ein variationsprinzip zur losung von Dirichlet-problemen bei verwendung von teilraumen, die keinen randbedingungen unterworfen sind. Abh Math Univ Hamburg 36:915
47. Hansbo P, Hermansson J (2004) Nitsches method for coupling
non-matching meshes in fluid-structure vibrational problems.
Comput Mech 32:134139
48. Arnold DN, Brezzi F, Cockburn B, Marini LD (2002) Unified
analysis of discontinuous Galerkin methods for elliptic problems.
SIAM J Num Anal 39:17491779
49. Bazilevs Y, Calo VM, Zhang Y, Hughes TJR (2006) Isogeometric fluidstructure interaction analysis with applications to arterial
blood flow. Comput Mech 38:310322
50. Bazilevs Y, Calo VM, Hughes TJR, Zhang Y (2008) Isogeometric
fluidstructure interaction: theory, algorithms, and computations.
Comput Mech 43:337
51. Bazilevs Y, Gohean JR, Hughes TJR, Moser RD, Zhang
Y (2009) Patient-specific isogeometric fluidstructure interaction
analysis of thoracic aortic blood flow due to implantation of the
Jarvik 2000 left ventricular assist device. Comput Methods Appl
Mech Eng 198:35343550
52. Bazilevs Y, Hsu M-C, Benson D, Sankaran S, Marsden
A (2009) Computational fluidstructure interaction: methods and
application to a total cavopulmonary connection. Comput Mech
45:7789
53. Bazilevs Y, Hsu MC, Zhang Y, Wang W, Liang X, Kvamsdal
T, Brekken R, Isaksen J (2010) A fully-coupled fluidstructure
interaction simulation of cerebral aneurysms. Comput Mech 46:
316
54. Bazilevs Y, Hsu M-C, Zhang Y, Wang W, Kvamsdal T, Hentschel S, Isaksen J (2010) Computational fluidstructure interaction: methods and application to cerebral aneurysms. Biomech
Model Mechanobiol 9:481498
55. Zhang Y, Wang W, Liang X, Bazilevs Y, Hsu M-C, Kvamsdal T,
Brekken R, Isaksen JG (2009) High-fidelity tetrahedral mesh generation from medical imaging data for fluidstructure interaction
analysis of cerebral aneurysms. Comput Model Eng Sci 42:131
150
56. Hsu M-C, Bazilevs Y (2011) Blood vessel tissue prestress modeling for vascular fluidstructure interaction simulations. Finite Elem
Anal Des 47:593599
57. Long CC, Hsu M-C, Bazilevs Y, Feinstein JA, Marsden
AL (2012) Fluidstructure interaction simulations of the Fontan
procedure using variable wall properties. Int J Num Methods Biomed Eng 28:512527
58. Brooks AN, Hughes TJR (1982) Streamline upwind/Petrov-Galerkin formulations for convection dominated flows with particular
emphasis on the incompressible NavierStokes equations. Comput Methods Appl Mech Eng 32:199259

123

Comput Mech (2012) 50:821833


59. Hughes TJR, Tezduyar TE (1984) Finite element methods for
first-order hyperbolic systems with particular emphasis on the
compressible Euler equations. Comput Methods Appl Mech Eng
45:217284
60. Tezduyar TE, Park YJ (1986) Discontinuity capturing finite element formulations for nonlinear convection-diffusion-reaction
equations. Comput Methods Appl Mech Eng 59:307325
61. Hughes TJR, Franca LP, Balestra M (1986) A new finite element
formulation for computational fluid dynamics: V. Circumventing
the BabukaBrezzi condition: a stable PetrovGalerkin formulation of the Stokes problem accommodating equal-order interpolations. Comput Methods Appl Mech Eng 59:8599
62. Tezduyar TE, Osawa Y (2000) Finite element stabilization parameters computed from element matrices and vectors. Comput Methods Appl Mech Eng 190:411430
63. Tezduyar TE (2003) Computation of moving boundaries and interfaces and stabilization parameters. Int J Num Methods Fluids
43:555575
64. Hughes TJR, Scovazzi G, Franca LP (2004) Multiscale and stabilized methods. In Stein E, de Borst R, Hughes TJR, (eds) Encyclopedia of computational mechanics, vol 3, Fluids ch 2. Wiley,
Weinheim
65. Tezduyar TE, Sathe S (2007) Modeling of fluidstructure interactions with the spacetime finite elements: solution techniques. Int
J Num Methods Fluids 54:855900
66. Bazilevs Y, Calo VM, Cottrel JA, Hughes TJR, Reali A, Scovazzi
G (2007) Variational multiscale residual-based turbulence modeling for large eddy simulation of incompressible flows. Comput
Methods Appl Mech Eng 197:173201
67. Hsu M-C, Bazilevs Y, Calo VM, Tezduyar TE, Hughes
TJR (2010) Improving stability of stabilized and multiscale formulations in flow simulations at small time steps. Comput Methods
Appl Mech Eng 199:828840
68. Bazilevs Y, Michler C, Calo VM, Hughes TJR (2007) Weak Dirichlet boundary conditions for wall-bounded turbulent flows.
Comput Methods Appl Mech Eng 196:48534862
69. Bazilevs Y, Michler C, Calo VM, Hughes TJR (2010) Isogeometric variational multiscale modeling of wall-bounded turbulent flows
with weakly enforced boundary conditions on unstretched meshes.
Comput Methods Appl Mech Eng 199:780790
70. Bazilevs Y, Akkerman I (2010) Large eddy simulation of turbulent
TaylorCouette flow using isogeometric analysis and the residual
based variational multiscale method. J Comput Phys 229:3402
3414
71. Bazilevs Y, Calo VM, Cottrell JA, Evans JA, Hughes TJR, Lipton
S, Scott MA, Sederberg TW (2010) Isogeometric analysis using
T-splines. Comput Methods Appl Mech Eng 199:229263
72. Dorfel MR, Juttler B, Simeon B (2010) Adaptive isogeometric
analysis by local h-refinement with T-splines. Comput Methods
Appl Mech Eng 199:264275
73. Cirak F, Ortiz M, Schroder P (2000) Subdivision surfaces: a new
paradigm for thin shell analysis. Int J Num Methods Eng 47:2039
2072
74. Cirak F, Ortiz M (2001) Fully C 1 -conforming subdivision elements for finite defomation thin shell analysis. Int J Num Methods
Eng 51:813833
75. Cirak F, Scott MJ, Antonsson EK, Ortiz M, Schroder P (2002) Integrated modeling, finite-element analysis, and engineering design
for thin-shell structures using subdivision. Comput Aided Des
34:137148
76. Onate E, Zarate F (2000) Rotation-free triangular plate and shell
elements. International Journal for Numerical Methods in Engineering 47:557603
77. Onate E, Flores FG (2005) Advances in the formulation of the rotation-free basic shell triangle. Comput Methods Appl Mech Eng
194:24062443

Comput Mech (2012) 50:821833


78. Benson DJ, Bazilevs Y, Hsu M-C, Hughes TJR (2011) A large
deformation, rotation-free, isogeometric shell. Comput Methods
Appl Mech Eng 200:13671378
79. Nguyen-Thanh N, Kiendl J, Nguyen-Xuan H, Wchner R, Bletzinger KU, Bazilevs Y, Rabczuk T (2011) Rotation-free isogeometric
thin shell analysis using PHT-splines. Comput Methods Appl Mech
Eng 200:34103424
80. Farhat C, Lesoinne M, Le Tallec P (1998) Load and motion transfer algorithms for fluid/structure interaction problems with nonmatching discrete interfaces: momentum and energy conservation,
optimal discretization and application to aeroelasticity. Comput
Methods Appl Mech Eng 157:95114
81. Takizawa K, Tezduyar TE (2011) Multiscale spacetime fluid
structure interaction techniques. Comput Mech 48:247267
82. Takizawa K Tezduyar TE (2012) Spacetime fluidstructure interaction methods. Math Models Methods Appl Sc (To appear).
doi:10.1142/S0218202512300013
83. Tezduyar TE, Sathe S, Stein K (2006) Solution techniques for the
fully-discretized equations in computation of fluidstructure interactions with the spacetime formulations. Comput Methods Appl
Mech Eng 195:57435753
84. Tezduyar TE, Sathe S, Keedy R, Stein K (2006) Spacetime finite
element techniques for computation of fluidstructure interactions.
Comput Methods Appl Mech Eng 195:20022027
85. Chung J, Hulbert GM (1993) A time integration algorithm for
structural dynamics withimproved numerical dissipation: the
generalized- method. J Appl Mech 60:371375
86. Jansen KE, Whiting CH, Hulbert GM (2000) A generalized-
method for integrating the filtered NavierStokes equations with
a stabilized finite element method. Comput Methods Appl Mech
Eng 190:305319
87. Tezduyar TE, Behr M, Mittal S, Johnson AA (1992) Computation
of unsteady incompressible flows with the finite element methods:
spacetime formulations, iterative strategies and massively parallel
implementations. In New methods in transient analysis. PVP-vol
246/AMD-vol 143, pp 724, ASME, New York

833
88. Tezduyar T, Aliabadi S, Behr M, Johnson A, Mittal S (1993) Parallel finite-element computation of 3D flows. Computer 26(10):27
36
89. Johnson AA, Tezduyar TE (1994) Mesh update strategies in parallel finite element computations of flow problems with moving
boundaries and interfaces. Comput Methods Appl Mechanics Eng
119:7394
90. Jonkman JM (2009) Dynamics of offshore floating wind turbines:
model development and verification. Wind Energy 12:459492
91. Jonkman JM, Matha D (2011) Dynamics of offshore floating wind
turbines: analysis of three concepts. Wind Energy 14:557569
92. Lackner MA, Rotea MA (2011) Passive structural control of offshore wind turbines. Wind Energy 14:373388
93. Nicklasson PJ, Homola MC, Virk MS, Sundsb PA (2012) Performance losses due to ice accretion for a 5 MW wind turbine. Wind
Energy 15:379389
94. Chow R, van Dam CP (2011) Verication of computational simulations of the NREL 5 MW rotor with a focus on inboard ow
separation. Wind Energy. doi:10.1002/we.529
95. Lackner MA (2012) An investigation of variable power collective
pitch control for load mitigation of floating offshore wind turbines.
Wind Energy. doi:10.1002/we.1500
96. Sebastian T, Lackner MA (2012) Characterization of the unsteady
aerodynamics of offshore floating wind turbines. Wind Energy.
doi:10.1002/we.545
97. Hsu M-C, Akkerman I, Bazilevs Y (2011) High-performance computing of wind turbine aerodynamics using isogeometric analysis.
Comput Fluids 49:93100
98. Hsu M-C (2012) FluidStructure Interaction Analysis of Wind Turbines PhD thesis. University of California, San Diego
99. Hau E (2006) Wind turbines: fundamentals, technologies, application, economics, 2nd edn. Springer, Berlin

123

You might also like