You are on page 1of 24

Renewable Energy 76 (2015) 338e361

Contents lists available at ScienceDirect

Renewable Energy
journal homepage: www.elsevier.com/locate/renene

Coupled multi-body dynamics and CFD for wind turbine simulation


including explicit wind turbulence
Y. Li a, A.M. Castro a, T. Sinokrot b, W. Prescott b, P.M. Carrica a, *
a
IIHR-Hydroscience and Engineering, The University of Iowa, Iowa City, IA 52242, USA
b
LMS International, a Siemens Business, 2425 Oakdale Boulevard, Coralville, IA 52241, USA

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

Article history: A high fidelity approach for wind turbine aero-elastic simulations including explicit representation of the
Received 19 November 2013 atmospheric wind turbulence is presented. The approach uses a dynamic overset computational fluid
Accepted 6 November 2014 dynamics (CFD) code for the aerodynamics coupled with a multi-body dynamics (MBD) code for the
Available online 2 December 2014
motion responses to the aerodynamic loads. Mann's wind turbulence model was implemented into the
CFD code as boundary and initial conditions. The wind turbulence model was validated by comparing the
Keywords:
theoretical one-point spectrum for the three components of the velocity fluctuations, and by comparing
Wind turbine aerodynamics
the expected statistics from the CFD simulated wind turbulent field with the explicit wind turbulence
Computational fluid dynamics
Multi-body dynamics
inlet boundary from Mann model. Extensive simulations based on the proposed coupled approach were
Wind turbulence conducted with the conceptual NREL 5-MW offshore wind turbine in an increasing level of complexity,
Wake flows analyzing the turbine behavior as elasticity, wind shear and atmospheric wind turbulence are added to
the simulations. Results are compared with the publicly available simulations results from OC3 partici-
pants, showing good agreement for the aerodynamic loads and blade tip deflections in time and fre-
quency domains. Wind turbulence/turbine interaction was examined for the wake flow. It was found that
explicit turbulence addition results in considerably increased wake diffusion. The coupled CFD/MBD
approach can be extended to include multibody models of the shaft, bearings, gearbox and generator,
resulting in a promising tool for wind turbine design under complex operational environments.
© 2014 Elsevier Ltd. All rights reserved.

1. Introduction speeds and less turbulence. However, other problems arise,


including waves and platform stability. The complex operational
Current trends of wind turbine design favor larger size which environment causes another critical factor as maintenance cost,
tends to be more cost-effective. However, this results in more especially for offshore farms with expensive accessibility. Of all
complex turbine systems with more demanding structural con- components in a turbine, gearbox, drivetrain and generator
straints. These large turbines, with rotor diameters over 120 m, contribute most to downtime, while rotor hub and blades are the
expose the wind turbine to high wind shear and turbulence. The next critical factors [1].
long and slender blades are subject to large amplitude changes in Development of methodologies and techniques capable of
wind loads, causing reliability issues due to fatigue. Variable-speed, modeling the interaction between realistic wind loads and the
variable-pitch and yaw control are needed for the turbine system to structural components is the most promising way to improve de-
achieve best performance. High tip speeds due to the long blades signs that will better perform in complex operational environ-
introduce noise and environmental impacts. Challenges increase ments. This covers the fields of aerodynamics, elasticity and wind
when several wind turbines are operated as wind farms, with simulations.
stronger velocity gradients and fluctuations caused by momentum Due to advances in high performance computing (HPC) tech-
deficits and wake turbulence of upwind turbines. The above- niques, the high fidelity but computationally expensive computa-
mentioned issues tend to be relieved when operating offshore, tional fluid dynamics (CFD) methods have shown great potential for
where the turbines are exposed to higher, more constant wind accurate wind turbine aerodynamic predictions. CFD has the
advantage over traditional, lower cost methods like Blade Element
* Corresponding author. Tel.: þ1 319 335 6381. Momentum (BEM) that all the geometry is resolved and the forces
E-mail address: pablo-carrica@uiowa.edu (P.M. Carrica). and moments over all the structure are predicted, and not subject

http://dx.doi.org/10.1016/j.renene.2014.11.014
0960-1481/© 2014 Elsevier Ltd. All rights reserved.
Y. Li et al. / Renewable Energy 76 (2015) 338e361 339

to external input of lift, drag and momentum coefficients. Also, directly by Direct Numerical Simulation (DNS). However, the
BEM-based models are greatly dependent on empirical corrections computational cost is way beyond current computer capabilities.
to account for 3D effects, unsteady inflow and dynamic stall, all of Another accurate but less expensive method is Large Eddy Simu-
these naturally predicted by CFD. Another advantage is that the lation (LES), which is an approximate solution to the NeS equations
wake and its turbulence are predicted, which is not possible with where the smallest scales are not solved directly but modeled [16].
BEM. These advantages come at a cost of orders of magnitude more One advanced atmospheric turbulent model was proposed by
expensive runs, but opens the possibility of better results for Bechmann [17]. Based on LES and incorporated in the CFD code
transient loads and wakes. EllipSys3D, the model provides improved representation of the
Various CFD or hybrid BEM/CFD methods have been attempted anisotropy of the atmospheric turbulence. However, LES still re-
for wind turbine simulations. Examples include the use of the quires unmanageably large computational resources and is not yet
generalized actuator disc method, combining BEM method and the practical for engineering use. Currently the most adopted models
NaviereStokes (NeS) equations, to represent the blade geometry for wind turbulence simulations are based on the construction of
by surface forces acting upon the incoming flow. This method has spectral tensors such as the Sandia model [18] and the Mann wind
been applied to study turbines and turbine wakes [2] and wind turbulence model [19,20]. Mann's model is capable of simulating all
farms [3]. The method was improved by the more sophisticated three velocity components of a 3D incompressible turbulence field.
actuator line [4] and actuator surface [5] approaches. Other Most of the aerodynamic codes capable of including wind turbu-
methods modeled the rotor directly by constructing body-fitted lence are based on BEM, and thus interaction between wind tur-
grids, including unstructured multi-grid RANS [6], compressible bines and the turbulent field cannot be predicted. In this respect,
RANS with overset grid technique [7] and NURB-based (Non-Uni- CFD provides a good methodology to simulate turbulence/turbine
form Rational B-splines) approach for turbine geometry [8]. interaction. Troldborg et al. [21] studied wind turbine wake with
There are two widely used approaches for structural modeling both resolved rotor geometry and rotor model using actuator line/
of modern wind turbines, finite element methods (FEM) and multi- actuator disc methods, applying Mann's model for the turbulent
body dynamics (MBD). FEM allows for complex blade deformations, inflow. The turbulent velocity field was introduced via body forces
which is common for large-scale wind turbines, but can be very in the momentum equation. Schulz et al. [22] conducted a DES
costly. The long and slender blade justifies the use of beam theory simulation for wind turbine in both flat and complex terrain to
in the structural modeling, with the classic beam theory for small evaluate the influence of inflow turbulence on blade load and po-
deformations or second-order non-linear beam theory for large wer response.
deformations. The dynamic interaction between components of the The objective of this paper is to present a high fidelity approach
turbine system in large-scale become important, including the for wind turbine aero-elastic simulation including the atmospheric
rotor-shaft-gearbox-generator dynamics. This involves different wind turbulence and wind shear. This is achieved by coupling an
body motions for each component of the turbine system, in which overset dynamics CFD solver to predict the aerodynamics of the
the components are combined with connections where loads and turbine in motion, and an MBD solver to predict the motions of the
displacements are communicated from one component to the other turbine system under the aerodynamic loads. The Mann model
[9]. MBD is highly accurate to model slender bodies [10] and much recommended in the IEC 61400-1 ed. 3 standard [23] is used to
faster than FEM, and is thus adopted by most of the simulation tools explicitly model turbulence, and was implemented into the CFD
for wind turbine analysis. code as inlet boundary and initial conditions. Extensive simulations
Coupling BEM and MBD perhaps is the most widely adopted based on the proposed approach were conducted with the NREL 5-
approach for wind turbine aero-elastic simulations to date, allow- MW offshore wind turbine. Results were compared and analyzed
ing efficient and good predictions for rotor aerodynamics and non- with the publicly available simulation results from offshore code
linear structural responses. These include the two primary design comparison collaboration (OC3) participants.
codes applied by the U.S. wind industry, AeroDyn/FAST and Aero-
Dyn/ADAMS. However, limits from the aerodynamic predictions
restrict the BEM/MBD applications. A more advanced aerodynamic 2. Mathematical and numerical methods
predictor such as CFD provides a better solution, with less modeled
inputs and the ability to predict the turbine wake. 2.1. Structural solver
Most efforts to couple CFD and computational structural dy-
namics (CSD) solvers have been reported in the rotorcraft com- The multi-body dynamics (MBD) simulation code Virtual.Lab
munity. Examples are the coupled overset CFD solver Overflow-D Motion [24,25], is used in simulating the structural dynamics for
and the flexible multi-body dynamics code DYMORE [11] and the turbine. Virtual.Lab Motion uses a set of generalized co-
coupled unsteady RANS unstructured grid solver FUN3D and the ordinates that are based on a generalized Cartesian coordinate
rotorcraft CSD code CAMRAD II [12]. CFD/CSD coupled studies for system (X, Y, Z) and Euler parameters (e0, e1, e2, e3) to formulate the
wind turbines are also available. Corson et al. [13] performed equations of motion [26]. The multi-body system consists of
coupled fluidestructure interaction (FSI) simulations for a 13.2 MW interconnected bodies that can be rigid or flexible, each of which
blade design using the commercial CFD solver AcuSolve and the may have translational and rotational displacements. The bodies
MBD solver FAST. Bazilevs et al. [14] proposed an FSI procedure for are connected by force and joint elements that describe their dy-
wind turbine simulations using a FEM based CFD solver and a namic and kinematic constraints. The code has the capability to
structural solver based on the isogeometric rotation-free Kircho- simulate realistic motions of complex mechanical system such as
hoffeLove composite shell and the bending strip method. The vehicles and powertrains.
coupled approach was applied to both rotor-only and the full tur- The motion of constrained bodies in the MBD code is described
bine configuration including tower and nacelle [15] for the NREL 5- by a set of differential-algebraic equations (DAEs) that consist of the
MW baseline wind turbine, showing a good combination of accu- differential equations of motion and a set of algebraic constraint
racy and efficiency. equations
A realistic transient turbulent wind field is important for wind
turbine simulations. In general, the most “correct” and accurate
fðq; tÞ ¼ 0 (1)
way to simulate the turbulent field is to solve the NeS equations
340 Y. Li et al. / Renewable Energy 76 (2015) 338e361

The dimensionless momentum and continuity equations for


i
€ þ fTq l ¼ Q iv þ Q ie
Miq (2) incompressible flows are
i

" #
where q ¼ ½q1 q2 … qn T is the vector of the generalized co- vu 1  
_
þ V$½ðu  xÞ5u ¼ Vp þ V$ Vu þ VuT
þS (5)
ordinates that consists of the translational and rotational co- vt Reeff
ordinates of each body in the system measured in the global frame
and n is the number of bodies in the system. f is a set of kinematic
V$u ¼ 0 (6)
constraints, fTqi is the Jacobian of the vector of constraints f with
respect to the generalized coordinates of body i, and l is the vector
where u is the fluid velocity, x_ is the grid velocity to account for
of the Lagrange multipliers for the constraints. Mi is the mass
moving or deforming grids, 5 denotes dyadic product, and S is a
matrix of the body i, Qiv is the quadratic velocity vector used to
source term due to body forces, e.g. a rotor or propeller model, and
describe Coriolis and centrifugal terms, obtained by the partial
is zero in this paper for explicit wind turbine simulations. p is the
derivatives of the kinetic energy of the body with respect to time t
dimensionless piezometric pressure, p p pabs/rU20 þ z/Fr2 þ 2k/3
¼ ffiffiffiffiffi
and the generalized coordinates q; Qie is the vector of the external
with pabs the absolute pressure, Fr ¼ U0 = gL is the Froude number;
forces applied on the body i.
Reeff is the effective Reynolds number, defined as Re1 eff ¼ Re
1
þ nt ,
In order to solve the set of DAEs (Eqs. (1) and (2)), Eq. (1) is
Re ¼ rlU0L/ml is the Reynolds number for water (l ¼ w) or air (l ¼ a). k
modified by taking the second partial derivative with respect to
and nt are the dimensionless turbulent kinetic energy and turbulent
time t
eddy viscosity, respectively, obtained from the turbulence model. In
  this work we use a delayed detached eddy simulation (DDES)
€ ¼ ftt  fq q_ q_  2fqt q_ ¼ g
fq q (3)
q model [33] based on Menter's shear stress transport model (SST)
[34], a two equation model for the turbulent kinetic energy k and
where subscript q and t denote their partial derivative, respectively.
the specific dissipation rate u. No boundary layer transition model
The DAEs can then be written in matrix form by combining Eqs.
was used, implying that the boundary layer is considered fully
(2) and (3) as
turbulent.
     It is noted that conservation laws (mass, momentum, geometry)
M fTq €
q Qv þ Qe
¼ (4) are satisfied inside the computational domain but not in cells using
fq 0 l g
fringe overset points. Conservative overset algorithms are difficult
to implement in general curvilinear grids [35], and overset con-
where Qe contains the external loads and is where the wind loads
servation errors are usually small if good practices are followed
are added.
when constructing the grids.

2.2. Flow solver 2.3. Coupling strategy

The incompressible CFD code CFDShip-Iowa v4.5 is used as the The coupled aero-elastic approach is done by exchanging the
flow solver for the turbine simulations. Its ability to accurately necessary information between the CFD and the MBD codes at run
simulate complex turbine aerodynamics has been extensively time. The forces and moments computed by the CFD code are sent
tested against data for the NREL phase VI wind turbine [27]. to the MBD code as Qie for each body in Eq. (2), and the positions/
CFDShip-Iowa v4.5 is a finite difference, unsteady Reynolds- rotations obtained by the MBD code are sent to the CFD code and
Averaged NaviereStokes (URANS) or Detached Eddy Simulation used to move the CFD grids to recompute the convective term and
(DES) solver. It uses structured multi-block body-fitted curvilinear solid boundary conditions.
grids with overset capabilities to accommodate complex geome- Fig. 1 schematically describes the approach for discretization.
tries and motions. Both single phase and two-phase (air/water) Blades and tower are slender and thus can be well approximated by
flow problems can be considered. The air/water problem is flexible one-dimensional structures. The system is discretized into
addressed by using a semi-coupled method [28] where the water interconnected bodies and represented in the MBD code by their
flow is decoupled from the air solution, but the air flow uses the generalized coordinates at the center of gravity (CG) with proper
unsteady water flow as a moving immersed boundary condition. position and orientation. The interconnection between bodies is
The free surface is modeled with an unsteady single-phase level set described and constrained by bracket joints for rigid turbine sim-
capturing approach [29], where only the water flow is solved with ulations that allow no motions between two bodies, and beam force
enforced kinematic and dynamic free surface boundary conditions elements for flexible turbine simulations that allow 6 DOF for each
on the interfaces, allowing robust computations and large ampli- body. Since the CFD grids are much finer than the discretization in
tude and/or steep waves. This approach allows for the simulation of the MBD code, special treatment was done to accommodate the
an off-shore wind turbine attached to a floating structure under information exchange for this non-matching domain discretization,
incoming waves, though the capability is not demonstrated here. as shown in Fig. 1. The CFD grids used to discretize the geometry
Dynamic overset grids are used to solve grid deformations and (blades, tower, etc.) consist of a body-fitted surface grid attached to
relative motions [30], with the interpolation coefficients between solid surfaces, and extend into the volume of the fluid to resolve the
the grids recomputed dynamically at run time with the code Suggar flow field. Each grid cell is assigned a set of integer IDs that are
[31]. Large-scale computations are achieved using high perfor- associated with the bodies it belongs to. As an example, a cell may
mance computing (HPC) with an MPI-based domain decomposition have ID values i and i  1 indicating that some part of the cell is
approach [32]. For complete details the reader is referred to the associated with body i and another part with body i  1. Integration
references [28e30,32] and literature therein. of the forces and moments for each cell on the surface grid is
All variables and properties are non-dimensionalized by the performed with respect to the CG of the corresponding body to
characteristic length L and velocity U0 of interest, and the liquid obtain the forces and moments contributing to each body, as
properties. For wind turbine simulations, they are chosen as the needed by the MBD code. To integrate the forces and moments
rotor radius and the prevailing incoming wind velocity. correctly, weights of the contribution for each cell to its associated
Y. Li et al. / Renewable Energy 76 (2015) 338e361 341

Fig. 1. Schematic demonstration of information exchange.

body are estimated based on the fraction of cell area belonging to both codes non-linear iterations are needed, due to the nonlinear
that body. The MBD code provides the motions for each body in nature of the equations of motion of the constrained multi-body
terms of 3 global positions and the rotational matrix with respect to system and the fluid flow. The CFD code obtains the overset
its initial configuration. Due to issues to be discussed later in this domain-connectivity-information (DCI) from Suggar at run time,
section, positions/rotations from more than one body will be and then non-linear iterations are performed to properly couple
applied to the cell even if the cell is only associated with one body. turbulence, level set, non-linear convection terms of the mo-
Fig. 2 shows a flow chart depicting the strategy to couple the mentum equations, and motions. Forces and moments are
CFD code and the MBD code. The coupling between structure and computed after the pressure implicit split operator (PISO) loop to
fluid is made in explicit form. This is adequate for the case of a wind obtain solenoidal velocity field. When the computations reach the
turbine since the added mass due to flow acceleration by the mo- communication time, positions/rotations are sent by the MBD code
tion of the structure is negligible compared to the mass of the to the CFD code which deforms the grids and obtains the new DCI
blades, tower and other moving components. for the next time step, while the updated forces and moments are
Communication and exchange of the forces and moments sent to the MBD code to compute the new motion responses. The
computed from the CFD code and the positions/rotations computed MBD code is much faster than the CFD code, thus MBD code shares
from the MBD code are needed so each program has the necessary one processor with the CFD code with no significant performance
information to perform its computations. Communication is made penalty.
through two communication files, one for forces and moments, and Some important issues need to be addressed for a successful
one for positions and rotations, so that the CFD code writes the coupling of the CFD and MBD codes. One is related to the imple-
appropriate forces and moments that the MBD code needs to mentation of the positions provided by the MBD code into the CFD
perform the multi-body dynamic computations, and the MBD code grids. Since the MBD model is composed by rigid bodies connected
writes the computed positions and rotations resulting from the by flexible beam connections, direct implementation of the body
input forces and moments. During the initialization stage, the CFD positions into the CFD grids can (and will) cause a collapse of the
code reads and splits the grids according to user directives for surface grids as the bodies partially overlap each other on moderate
parallel decomposition, while the MBD code reads the information or large deformations. To prevent this unphysical behavior, the new
of the system model that includes mass, position and orientation of position of the surface grid point is obtained by a weighted average
each body, structural properties and kinematic constraints of the of the positions of the two bodies whose CGs bound the grid point.
system's components. Then the CFD code sends the initial forces Since the CGs behave physically and do not collapse, the new grid
and moments to the MBD code, zero for a simulation starting from obtained with this interpolation process is well behaved.
scratch or the values from a specific restart solution for a restart Another issue related to the CFD grid deformation and motion is
run. Afterwards, the two codes begin their independent computa- related to refinement grids, designed to capture flow features of
tions until the time for communication specified by the user. For importance like tip vortices and regions of separation. These grids
342 Y. Li et al. / Renewable Energy 76 (2015) 338e361

Fig. 2. Coupling strategy.

have to follow the blades accurately, which for deforming blades with * denoting conjugation and 〈,〉 ensemble averaging. Eq. (8) is
becomes difficult. An average of the motions of all bodies belonging valid for infinitesimally small dki. For isotropic flows the velocity
to each blade is used to move the refinement grids attached to the spectrum tensor Fij is related to the three dimensional energy
blade. spectrum by

2.4. Wind turbulence model


EðkÞ  2 
Fij ðkÞ ¼ d ij k  ki kj (9)
The wind turbulence model proposed and developed by Mann 4pk4
[19,20] was implemented into the code CFDShip-Iowa v4.5 to
provide appropriate initial and inlet boundary conditions. Based on where the energy spectrum E(k) is modeled using the von K n
arma
the construction of a velocity-spectrum tensor for atmospheric model
surface-layer turbulence, the model includes turbulence fluctua-
tions and the effect of shear in the atmospheric boundary layer. By
construction, the model reproduces the second-order statistics as
L4 k4
found in the atmosphere and provides a solenoidal velocity field. EðkÞ ¼ a32=3 L5=3  17=6 (10)
Particular to this work is the computation of a valid pressure field 1 þ L2 k2
consistent with the generated velocity that is used to impose
appropriate inlet and initial conditions. with a z 1.5 the Kolmogorov constant, 3 the turbulence dissipation
and L a length scale that for high Reynolds numbers approaches the
limit L z 0.43L11 with L11 the integral length scale. The integral over
2.4.1. Spectral representation of the velocity field all wave numbers of the energy spectrum equals the total turbulent
The velocity is modeled as a stochastic field using the general- kinetic energy k. Performing this integration leads to a32/3L5/
3
ized FouriereStieltjes integral ¼ 1.453s2isoL which allows to eliminate the turbulence dissipation
Z by the mean turbulent fluctuations siso related to the turbulent
uðxÞ ¼ eik$x dZðkÞ (7) kinetic energy by k ¼ 3/2s2iso.
It is practically impossible to determine E(k) experimentally.
However, one-dimensional spectra can be determined by single-
where the integration in the wave number vector k spans the entire
point velocity measurements. A relationship exists between the
wave number space, and Z is a orthogonal stochastic process
one-dimensional spectra and the three dimensional spectrum [36]
related to the velocity-spectrum tensor Fij by
which if applied to the von Ka rman model in Eq. (10) allows to
obtain the one-dimensional spectra for the streamwise and trans-
〈dZi* ðkÞdZj ðkÞ〉 ¼ Fij ðkÞdk1 dk2 dk3 (8)
verse directions as
Y. Li et al. / Renewable Energy 76 (2015) 338e361 343

2.4.3. Computation of the pressure field


9 2=3 5=3 1 It is usually sufficient to have the velocity field as the solely wind
F1 ðk1 Þ ¼ a3 L  5=6 (11)
55 turbulence input for most of the simulation tools, and thus there is
1 þ ðLk1 Þ2
no description for the computation of the corresponding pressure
field in Mann model. However, the pressure field is needed for CFD
3 3 þ 8ðLk1 Þ2 simulations. This can be obtained by solving the NeS equations in
Fi ðk1 Þ ¼ a32=3 L5=3  11=6 ði ¼ 2; 3Þ (12) the frequency domain as done in the context of DNS simulations
110
1 þ ðLk1 Þ2 using spectral methods [36]. In the frequency domain pressure
modes are computed as.

b
kj G j
b
p ðkÞ ¼ i (20)
2.4.2. Discrete representation in Fourier modes k2
Eq. (7) is approximated by a discrete Fourier series as
b ¼ ik F u u
G (21)
X X j k j k
ui ðxÞ ¼ b ðkÞ ¼
eik$x u eik$x Cij ðkÞnj ðkÞ (13)
k k where F f$g denotes the Fourier transform operator and the hat
transformed variables. The product ujuk in Eq. (21) is computed in
where ub ðkÞ are the Fourier modes of the velocity field and nj(k) are
the physical domain with the generated velocity field
independent Gaussian complex variables with unit variance. In and then transformed to the frequency domain where Eqs. (20) and
Ref. [20] the relationship between the velocity spectrum and the (21) are evaluated. Pressure in the physical domain is then
coefficients Cij(k) is found leading to obtained by applying the inverse Fourier transform to Eq. (20) using
2 3 an FFT.
k2 z1 k3 þ k1 b  k1 z1 k2
!1=2 6 7
Eðk0 Þ$2p2 6 k3  k1 b þ k2 z2 k1 z2 k1 7
Cij ðkÞ ¼ 6 7
L1 L2 L3 k40 6 7
4 k20 k20 5 2.4.4. Implementation into CFDShip-Iowa v4.5
2
k2  2
k1 0 The general schematic for the implementation of Mann wind
k k
turbulence model into CFDShip-Iowa v4.5 as wind turbulence
(14) boundary condition is shown in Fig. 3, which also shows cross
sections in the Mann's box with non-dimensional axial velocity
z1 ¼ C1  k2 C2 =k1 ; z2 ¼ k2 C1 =k1 þ C2 (15) U. A stationary Mann wind turbulence box is generated as a pre-
processing step computing FFTs using the FFTW library [37].
 
bk21 k21 þ k22  k3 ðk3 þ bk1 Þ Since by construction the velocity field is periodic, the di-
C1 ¼  2  ; mensions of the box are chosen to be several integral length
k2 k1 þ k22
0 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1 (16)
scales such that velocities at the opposite faces of the box can be
k2 k20 bk1 k21 þ k22 considered to be uncorrelated. As depicted in Fig. 3, at time t ¼ 0
C2 ¼  3=2 arctan@ 2 A the front face of the Mann box is coincident with the inlet of the
k2 þ k2 k0  bk1 ðk3 þ bk1 Þ
1 2 CFD domain. Using Taylor's hypothesis of frozen turbulence, ve-
locity and pressure at the inlet are interpolated from a plane
G within the Mann box located at x ¼ L1  Uht, where L1 is the
b¼ sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi (17) longitudinal dimension of the Mann's box and Uh is the mean
2
k$ 1; 17; 4; k2
2 F1
3
3 6 3 wind velocity at hub height. This plane is made to go back to the
front of the box once it reaches the back using the fact that the
generated velocity field is periodic. This procedure is equivalent
where Li are the dimensions of the physical domain, to having the Mann box moving forward as depicted in Fig. 3.
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
k¼ k21 þ k22 þ k23 is the magnitude of the wave number and Using the same Taylor's hypothesis velocity and pressure are
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi initialized at t ¼ 0.
k0 ¼ k2 þ 2bk1 k3 þ ðbk1 Þ2 is the magnitude before shear distor-
tion. The model includes the effect of shear through the dimen-
sionless distortion time b where G ¼ 3.9 is used as recommended by 3. Simulation design
the IEC and 2 F1 is the hypergeometric function. The parameters of
the model siso and L are estimated following the IEC standard from 3.1. Simulation turbine
where
The conceptual NREL 5-MW offshore wind turbine [38] is used
0:56zh zh < 60 m as geometry for the simulations. Fig. 4 and Table 1 show the ge-
L½m ¼ (18)
33:6 m zh > 60 m ometry and basic properties of the turbine that consists of the
baseline rotor, nacelle, tower and floating support platform struc-
siso ½m=s ¼ 0:55Iref ð0:75Uh ½m=s þ 5:6Þ (19) ture. It is a utility-scale, conventional three-bladed upwind
variable-speed variable-pitch controlled turbine, and has been
where zh and Uh are the hub height and mean wind velocity at the widely used as the reference turbine by other researchers and wind
hub height, respectively. Units are indicated in brackets. Iref is a turbine industries, including the Offshore Code Comparison
reference turbulent intensity specified according to the wind clas- Collaboration project (OC3) and its continuation (OC4). Notice that
ses defined in the IEC standard. Once the Fourier modes u b ðkÞ are the design rated tip speed of the turbine is 80 m/s, with Mach
computed the physical velocity is obtained by inversion of Eq. (13) number less than 0.3, justifying the use of an incompressible code
using an FFT. as flow solver.
344 Y. Li et al. / Renewable Energy 76 (2015) 338e361

Fig. 3. Schematic of implementation for Mann wind turbulence model.

3.2. Grid design the tower grid form the body rotor-nacelle-tower, such that turbine
yaw motion can be considered. In addition, a child could be added
Fig. 5 shows the grid design for CFD simulations. The system to the body rotor such that the blade pitch motion can be controlled
consists of the grids defining the turbine, including accurate or predicted. The dynamic overset technique is applied to re-
geometrical representations of the three blades, tips and tower/ compute the overset coefficients at run time with the code Sug-
floating platform as documented in Refs. [38,39], and an approxi- gar. With this approach, full control or prediction of the turbine can
mate nacelle and hub due to insufficient geometric information for be realized with the rotating rotor, blade pitch and yaw control, and
these components. In addition to the turbine grids, a Cartesian motions of the floating turbine under winds and waves. For all
background grid is used to set for the boundary conditions, with the simulations in this paper a second-order implicit Euler scheme was
grids refined near the expected free surface location with the still used for temporal terms. For the spatial discretization, the
water level at z ¼ 0 m so that ocean wave motions (to be introduced convective terms are discretized with a fourth-order upwind biased
in a future work) can be well captured. The background grid scheme, while a second-order centered scheme is used for the
extends 5.1  x/R  5.1, 2.5  y/R  2.5, and 5.1  z/R  6 with viscous terms. The time step was chosen such that the blades rotate
the center of the turbine located at x/R ¼ 0 in order to minimize 0.5 degrees per time step. In total 20 rotor revolutions were
boundary effects. With this configuration, the presence of the tur- included for each simulation case.
bine does not significantly disturb the inlet flow. Grid spacing in the
air side is designed to be less than that of the Mann's box so as to
capture the wind turbulence generated by the explicit wind tur- 3.3. Structural model
bulence model. Three Cartesian blade refinement grids are used to
resolve the flow around the blades, and one Cartesian air refine- The multi-body system model for the turbine consists of blades,
ment grid is constructed to resolve the wake flow close to and hub, nacelle and tower/floating platform as the substructures or
behind the rotor. A Total of 14 grids are used with about 6 M grid components that are linked by the appropriate kinematical con-
points. Gaps are present between blade roots and hub and between straints at their interfaces, as can be seen in Fig. 6. Each turbine
hub and nacelle so that blade pitch and rotor shaft rotation are component comprises one or several rigid bodies connected by the
possible. Boundary conditions on solid surfaces are set as no-slip, relevant connection. The structural properties of each component
with grid spacing set so that the condition yþ  1 is satisfied as include mass and center of gravity, flap-wise, edge-wise, torsional
required by the turbulence model. Inlet condition for the back- and extensional section stiffness, as well as flap-wise and edge-wise
ground grid is specified with prescribed flow velocities of uniform, section inertia. In this paper, construction of the model follows the
log-law profile or turbulent winds, while zero-gradient boundary structural information specified in Refs. [38,39]. The structural
conditions are applied elsewhere with the exception of the exit, model of the turbine consists of 6 components: 3 blades, nacelle,
which uses zero second derivative condition on the velocity (zero hub, and tower/floating platform. Each blade component comprises
traction). Note that static water and fixed tower/support platform of 48 bodies, the tower/floating platform component has 11 bodies,
are considered in this paper, resulting in a no-slip boundary at the and the nacelle and hub only contain 1 body each. In total the tur-
free surface of the background grid. This uses the implementation bine multi-body system has 157 bodies. Bracket joints are used as
described in Ref. [28] to impose the free surface as an immersed the connection in all components for the rigid turbine simulations
boundary, with no roughness or wall functions. to prevent any relative motion between two bodies, while beam
The grids are organized in a parent/child hierarchy, as shown in force elements are used as connections for the flexible turbine
Table 2. The blades, tips, hub and the blade refinements together simulations, allowing 6 degrees of freedom (6 DOF) for each body.
form the body rotor that rotates around the shaft; nacelle, tower- For the kinematic description between turbine components,
platform, air refinement and the body rotor form the body rotor- appropriate kinetic joints or constraints are applied: each interface
nacelle, allowing motions of the whole turbine under winds and of the blade and hub is connected by the bracket joint, restraining
waves. Though not included in the present simulations, another relative motions between the components; a revolute joint is used
hierarchy can be constructed with body rotor and the nacelle grid between the hub and nacelle along the rotational axis of the rotor to
forming the body rotor-nacelle while the body rotor-nacelle plus allow rotor rotation and constrain other DOFs; similarly, another
Y. Li et al. / Renewable Energy 76 (2015) 338e361 345

Fig. 4. NREL 5-MW offshore wind turbine configuration.

revolute joint is used at the interface of nacelle and tower for the AeroDyn [41] as aerodynamic solver but different MBD-based
yaw motion. structural solvers, FAST [42] and Bladed Multibody were
compared (Hereby called NREL FAST and GH Bladed). Table 3 sum-
3.4. Simulation cases marizes all cases evaluated, where rigid turbine simulations are
named 2.x and flexible turbine simulations are labeled 3.x. The
The test cases were chosen with incremental level of complexity simulation matrix includes cases with moderate wind condition
from the publicly available OC3 results [40], to include rigid and with the mean hub height wind speed of 8 m/s and rotor speed of 9
flexible turbines with or without wind turbulence. Several re- RPM, rated wind condition with mean hub height wind speed of
visions for each simulation case performed independently by a 11.4 m/s and rotor speed of 12.1 RPM, uniformly distributed wind
group of international participants from universities, research in- profile, wind shear with log-law wind profile, wind turbulence, rigid
stitutions and industries with expertise in wind energy makes the and flexible turbine. For the wind turbulence, a box of
OC3 project a good benchmark for the utility-scale offshore wind 256  128  128 points was used with length increment DL ¼ 10 m
turbine. for Mann's model, generating a turbulence field with dimension of
The simulation matrix from OC3 phase I was simulated, 2560 m  1280 m  1280 m. Reference turbulence intensity is 0.14, a
excluding the water/wave effects. Simulations results of two par- medium turbulence level and is consistent with the OC3 simulation
ticipants from OC3 that both use the BEM-based aerodynamic code conditions.
346 Y. Li et al. / Renewable Energy 76 (2015) 338e361

Table 1 Table 2
Basic properties of NREL 5-MW offshore wind turbine. Grid system information.

Baseline turbine properties Name i max j max k max Total points Hierarchy
Rating 5 MW
Nacelle 51 48 61 149 K Rotor-nacelle
Rotor orientation, configuration Upwind, 3 blades
Tower 201 48 61 588 K Rotor-nacelle
Rotor diameter, hub diameter 126 m, 3 m
RefAir 48 51 61 149 K Rotor-nacelle
Hub height 90 m
Hub 51 48 61 149 K Rotor
Cut-in, rated, cut-out wind speed 3 m/s, 11.4 m/s, 25 m/s
Blade1 151 48 61 442 K Rotor
Cut-in, rated rotor speed 6.9 RPM, 12.1 RPM
Tip 1 51 48 61 149 K Rotor
Rated tip speed 80 m/s
Blade2 151 48 61 442 K Rotor
Control Variable speed, collective pitch
Tip 2 51 48 61 149 K Rotor
Drivetrain High speed, multiple-stage gearbox
Blade 3 151 48 61 442 K Rotor
Overhang, shaft tilt, precone 5 m, 5 , 2.5
Tip 3 51 48 61 149 K Rotor
Blade Ref 67 48 92 295 K Rotor
Tower properties Blade Ref 67 48 92 295 K Rotor
Elevation to tower base above SWL 10 m Blade Ref 67 48 92 295 K Rotor
Elevation to tower top 87.6 m Background 244 101 123 3.03 M Earth
Total 6.7 M
Floating platform properties
Depth to platform base below SWL 120 m
Elevation to platform top above SWL 10 m While all simulation in this paper are performed at constant
Depth to top of taper below SWL 4m rotational speed, only case 2.1a in OC3 involves a constant rotor
Depth to bottom of taper below SWL 12 m
speed, as shown in Tables 3 and 4. The simulation case with con-
Platform diameter above taper 6.5 m
Platform diameter below taper 9.4 m stant rotor speed of 9 RPM at wind speed of 8 m/s for rigid turbine

Fig. 5. Grid system. Slice in x shows background (black) and wake refinement (green) grids. Axis-aligned slices on the blades show blade refinements (brown). Tip grid in (a) and a
horizontal cross-section showing overset grid topology around airfoil (b). (Grid points skipped in all directions for clarity; points skipped not the same for different grids.) (For
interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)
Y. Li et al. / Renewable Energy 76 (2015) 338e361 347

Fig. 6. Multi-body system model.

was to serve as the basic and simplest validation case in OC3 wind speed of 8 m/s, where NREL FAST determined that a rotor
project, and complexity was gradually added to include the turbine speed of 9.3 RPM was optimal while GH Bladed obtained 9.1 RPM.
control system, flexibility and wind turbulence. For wind speed Detailed discussion is included in the following sections.
below the rated wind speed of 11.4 m/s, variable rotor speed
controller with fixed blade pitch was applied to maximize turbine 4. Results and discussion
power, and variable pitch controller with fixed rotor speed was
applied for wind speeds beyond the rated wind speed so as to 4.1. Mann wind turbulence model predictions
maintain a constant rated power and regulate the generator speed.
One example can be found by comparing cases 2.1a and 2.1b at The validation of the explicit turbulence model is conducted in
two steps: first the generated Mann box is validated by comparing
with the theoretical isotropic one-point spectrum; then in step 2 a
Table 3 CFD wind turbulence simulation for an open field (i.e. no wind
Simulation conditions for the selected cases from OC3 phase I.
turbine) is conducted with the applied Mann wind turbulence
Case DOFs Wind condition Turbine model as inlet boundary and initial conditions. Statistical compar-
flexibility ison is made to show the validity of the implementation. To
2.1a None: constant rotor speed compare the isotropic one-point spectrum, the parameter (G) to
Steady, uniform,
and fixed blade pitch control the anisotropic wind fluctuations is set to 0 in step 1, and
no shear: Vhub ¼ 8 m/s
2.1b None: controlled rotor speed Rigid
thus the model reduces to isotropic wind turbulence. In addition,
2.2 None: controlled blade pitch Vhub ¼ 11.4 m/s, Iref ¼ 0.14 (B)
Turbulence model ¼ Mann the mean prevailing velocity is set to 0 for convenient comparison.
3.1 Tower, drivetrain, blades Steady, uniform, Flexible In step 2, all parameters were set appropriately to include aniso-
controlled rotor speed no shear: Vhub ¼ 8 m/s tropic and sheared wind turbulence, while mean velocity of 8 m/s
3.2 Tower, drivetrain, blades Vhub ¼ 11.4 m/s, Iref ¼ 0.14 (B) was used in the prevailing wind direction, resulting in a total
controlled blade pitch Turbulence model ¼ Mann
simulation time of 320 s. The CFD grid for this simulation consists
348 Y. Li et al. / Renewable Energy 76 (2015) 338e361

Table 4
Summary of OC3 results for wind speed 8 m/s.

Participant NREL FAST GH Bladed

Case 2.1a 2.1b 3.1 2.1a 2.1b 3.1

Rotor speed [RPM] Mean 9.000 9.334 9.329 9.000 9.136 9.128
Min 9.000 9.332 9.324 9.000 9.133 9.125
Max 9.000 9.337 9.335 9.000 9.139 9.132
s 0.000 0.002 0.003 0.000 0.002 0.002
Thrust [kN] Mean 384.97 394.74 409.63 372.32 375.62 378.61
Min 377.70 387.30 406.30 361.06 364.19 376.07
Max 387.00 396.90 412.70 374.49 377.90 381.26
s 2.60 2.66 1.46 3.77 3.82 1.36
Shaft torque [kN m] Mean 2096.40 2033.76 2031.62 1975.54 1948.36 1945.01
Min 2019.00 2025.00 2019.00 1863.94 1931.70 1933.54
Max 2115.00 2036.00 2043.00 1998.38 1958.56 1954.30
s 27.37 2.88 5.87 37.40 5.87 5.31
Power [kW] Mean 1975.81 1987.95 1984.80 1861.90 1864.04 1859.23
Min 1902.86 1979.34 1971.58 1756.72 1847.71 1847.86
Max 1993.34 1990.74 1996.72 1883.43 1873.13 1867.93
s 25.79 2.86 5.81 35.25 5.66 5.12

only of the same background grid used for turbine simulations. The the spectra for the three velocity components from the Mann
Mann model parameters used in step 2 are later applied to turbine model compare well with the theoretical curves. Due to the finite
simulations in turbulent wind. size of the Mann box and the cell, there are truncations in the
Fig. 7 shows a Mann's turbulence box with slices colored the calculated spectrum for low wave number (limited by the size of
axial velocity. The distribution of the velocity fluctuations is the box) and for large wave number (limited by the size of the
random in space, with velocities mostly within ±1 m/s and the cell). The wave number is bound by {2p/L1, 2p/DL} where L1 is the
largest wind fluctuations reaching ±3 m/s sparsely distributed in longitudinal dimension of the Mann box, and DL ¼ 10 m is its
the box. Fig. 8(a) shows the time history of velocities fluctuations length increment, leading to a resolvable range of wave numbers
at the point that corresponds to the position of the turbine hub. {0.0025 m1, 0.6283 m1} for the simulations.
From Taylor's frozen hypothesis, the spatial information in the Since the turbulence is introduced as initial and boundary
axial direction x is related to temporal information, and thus ve- conditions, there is a decay of turbulence in time, seen as a decay
locities at a point in the box moving in the x-direction is inter- with position in the CFD computation. Decay of the turbulence is
preted as velocities at a fixed point in space over time t. Since the due to the fact that no production such as shear exists downstream
turbulence is isotropic with zero mean velocity, all other points to balance the dissipation, with the only exception being the vi-
should have similar behavior. As can be seen in Fig. 8(a), all ve- cinity of the free surface where no-slip boundary exists due to the
locities fluctuate within ±3 m/s with zero net velocity, but no still water considered in the current simulations. This expected
similar or periodic pattern exists for the three velocity compo- phenomenon in a viscous computation has been observed by
nents in time. Statistics for all points in the box show zero mean Larsen [44], and is alleviated using a turbulence scaling factor (SF)
velocity and unit variance. In addition, Fig. 8(b) shows a compar- in the Mann model, which is calculated based on the actual vari-
ison of the one-point spectrum computed using the parametric ance level in the box and the target longitudinal turbulence stan-
spectral estimation method by Burg [43] and the theoretical one- dard deviation s1,target ¼ siso/0.55 based on the requested
point spectrum as defined in Eqs. (11) and (12). It can be seen that turbulence intensity, such that SF ¼ s1,target/s1,box multiplies all
velocity fluctuations. Fig. 9 shows the time history of turbulent
velocities for 4 points located at approximately hub height and
center of the turbine (y ¼ 0 m and z ¼ 90 m) and at different non-
dimensional axial locations including the inlet of the simulation
domain, right in the position of the turbine at x/R ¼ 0 and further
downstream to the exit of simulation domain approximately at x/
R ¼ 5. Recall that all spatial dimensions are non-dimensionalized by
the rotor radius R. It can be seen that for all locations the 3 velocities
components fluctuate randomly around their expected mean ve-
locities, 1 for u and 0 for v and w. The figure essentially shows that
the turbulent field solved by CFD with the explicit wind turbulence
at the inlet is appropriately transported through the whole simu-
lation domain throughout the simulation, but turbulence decay due
to viscous losses mentioned previously is observed, as velocities
closer to the inlet tend to have larger fluctuations than more
downstream locations.
Fig. 10 shows mean and standard deviation of the velocities on
vertical and horizontal transects crossing the turbine hub. The
mean velocity is maintained in u ¼ (1, 0, 0) while the standard
deviation for u (s1) satisfies the requested reference turbulent
intensity of 0.14 for axial velocity, and standard deviations for v
Fig. 7. Demonstration of the Mann turbulence box with axial velocity fluctuations. (s2) and w (s3) are about 0.1 and 0.08, respectively, satisfying the
Y. Li et al. / Renewable Energy 76 (2015) 338e361 349

Fig. 8. Mann wind turbulence box validation (a): velocity fluctuations along the streamwise direction; (b) one-dimensional spectrum for each velocity component computed from
the generated turbulence field compared against the theoretical (input) model spectrum.

requirements by the IEC standard that s2  0.7s1 and s3  0.5s1. not converged. The largest structures resolved are 400 m in
Note also that, due to the long time it takes for the largest tur- size (wave number 0.0025 m1 as stated earlier in this section,
bulent structures to cross the averaging points, the total averaging the dimensional velocity is 8 m/s, and the integration time is
time only covers 6.4 of the largest structures and the statistics are 320 s).

Fig. 9. Time history of turbulent velocities at hub height (y ¼ 0; z ¼ 90 m) and different axial positions.
350 Y. Li et al. / Renewable Energy 76 (2015) 338e361

Fig. 10. Flow field statistics for several vertical lines at x ¼ 0.

4.2. Aerodynamic predictions the time history of thrust and torque once the periodic behavior has
been reached for wind speed of 8 m/s, and Table 5 quantitatively
Case 2.1a provides a good scenario to compare results of the CFD compares statistics. As can be seen in Fig. 11, both CFD and the OC3
approach presented herein (from now on called the CFD approach results predict similar trends for thrust and torque for the rigid
or simply CFD) against the widely used aerodynamic solver Aero- turbine under uniformly distributed wind. Being a 3-bladed tur-
Dyn under fixed operational conditions without controller. For all bine, a decrease in thrust and torque occurs every 1/3 rotation due
performed simulations with CFD, at the initial time step blade 1 to the presence of the tower. Using GH Bladed as baseline, quan-
was placed downward right in front of the tower, and thus at every titative comparisons in Table 5 show that the CFD approach has
complete rotation blade 1 is passing the tower while half rotation close predictions to both NREL FAST and GH Bladed. For thrust, CFD
later is at the uppermost position. Turbine rotation is counter clock- predicts an average value of 389 kN with standard deviation of
wise when seen facing downwind. Fig. 11 shows a comparison of 2.7 kN, 4.5% larger than GH Bladed, while NREL FAST predicts

Fig. 11. Thrust and torque for wind speed 8 m/s.


Y. Li et al. / Renewable Energy 76 (2015) 338e361 351

Table 5
Thrust and torque for wind speed 8 m/s.

Participant CFD NREL FAST GH Bladed

Case Uniform rigid Log-law rigid 2.1a 2.1a

Thrust [kN] Mean/difference 388.9 (4.46%) 385.0 (3.41%) 385.0 (3.41%) 372.3 (0.00%)
Min/difference 382.0 (5.79%) 378.2 (4.74%) 377.7 (4.60%) 361.1 (0.00%)
Max/difference 392.5 (4.81%) 388.2 (3.66%) 387.0 (3.34%) 374.5 (0.00%)
s 2.71 2.75 2.60 3.77
Shaft torque [kN m] Mean/difference 1945.7 (1.51%) 1899.6 (3.84%) 2096.4 (6.12%) 1975.5 (0.00%)
Min/difference 1869.8 (0.32%) 1830.5 (1.79%) 2019.0 (8.32%) 1863.9 (0.00%)
Max/difference 1980.1 (0.92%) 1929.0 (3.47%) 2115.0 (5.83%) 1998.4 (0.00%)
s 27.53 26.53 27.37 37.40

Case Uniform, flex Log law, flex 3.1 3.1

Thrust [kN] Mean/difference 402.6 (6.34%) 398.2 (5.18%) 409.6 (8.19%) 378.6 (0.00%)
Min/difference 396.0 (5.29%) 391.4 (4.07%) 406.3 (8.03%) 376.1 (0.00%)
Max/difference 406.7 (6.66%) 401.7 (5.35%) 412.7 (8.23%) 381.3 (0.00%)
s 2.30 2.40 1.46 1.36
Shaft torque [kN m] Mean/difference 1934.6 (0.53%) 1886.1 (3.03%) 2031.6 (4.45%) 1945.0 (0.00%)
Min/difference 1853.3 (4.15%) 1811.2 (6.33%) 2019.0 (4.42%) 1933.5 (0.00%)
Max/difference 1974.5 (1.03%) 1918.4 (1.84%) 2043.0 (4.54%) 1954.3 (0.00%)
s 25.73 23.71 5.87 5.31

average thrust of 385 kN with standard deviation of 2.6 kN, 3.4%


larger than GH Bladed. For torque, CFD shows an average magni- aerodynamic blade-load calculations” with “differences in the
tude of 1946 kN m with standard deviation of 27.5 kN m, while mean magnitude of rotor torque of about 5%” for all participants.
NREL FAST predicts an average of 2096 kN m with standard devi- Regarding this fundamental test, CFD shows good agreement with
ation of 27.4 kN m, 1.5% lower and 6.1% higher than GH Bladed, both methods.
respectively. Beside the good agreement for averaged magnitudes, Further investigation of the individual blade loads is beneficial
all three methods exhibit similar statistics for this case, including to explain and understand the aerodynamic behavior of the rotor
maximum, minimum and standard deviation of the thrust and and quantify the effects of the tower shadow and tilt/precone angle.
torque. Notice that NREL FAST and GH Bladed share the same Fig. 12 shows the time history of thrust and torque for blade 1. The
aerodynamic code, and yet a difference of 3.4% is reported for thrust loads experienced by the blades show periodic oscillations, with a
and even a larger 6.1% difference for torque. As pointed out in the sharp drop of 6.3% in thrust and 10.6% in torque every time the
OC3 final report [40] “certain differences were apparent in the blade passes by the tower. The tilt causes a lower relative velocity of

Fig. 12. Thrust and torque for blade 1.


352 Y. Li et al. / Renewable Energy 76 (2015) 338e361

the wind with respect to the blade when the blade is rotating from transformed to the blade system as seen by the airfoil section. The
bottom to top, and the opposite when the blade is coming back AOA is defined as the difference in angle between the chord of the
down from top to bottom. This effect is caused by the tip of the airfoil and the flow direction far upstream of the airfoil. Use of the
blades moving back and forth in the axial direction as they rotate. velocity far upstream of the airfoil would be inappropriate since it
The tilt also results in a change of the angle of attack of airfoil would not contain the induced flow in the axial and tangential
sections, which increase as the blade moves upwards and decreases directions. Using the velocity too close to the airfoil does not work
when blade moves downwards, changing the sectional behavior either, since the influence of the airfoil affects the streamlines and
including axial and tangential forces. The consequence is fluctua- causes errors in the AOA. Necessarily, a compromise needs to be
tions in thrust and torque with approximately 2% and 5% amplitude, made to estimate the flow direction so that represents the flow
respectively. impacting the airfoil but not overly affected by it. Similarly, the
induction factors are computed from velocities obtained from CFD
4.3. Effect of elasticity and the location of the points selected to compute the velocities
used to estimate the induction factors affect the results. A more
Case 3.1 was designed to test aero-elastic capabilities of the advanced method to compute the AOA is described in Ref. [45] and
simulation codes by OC3. However, as mentioned in the simulation used here, but still results are not ideal. In light of this, the AOA and
cases section, OC3 simulations have a variable-speed variable-pitch the induction factors have to be considered a qualitative estimate to
controller, which is not included in the CFD simulations. An analysis evaluate behavior and trends.
is conducted here to estimate the influence of the controller before The average axial velocity over a set of points approximate half
comparing results with elasticity. The control algorithm seeks the chord length around the foil is used to compute the axial induction
maximal turbine power, with the rotor speed adjusted at run time factor a. Of those points, only the points on the suction side are used
and resulted in around 9.3 RPM for both the rigid and flexible to compute the tangential induction factor a0 . The force coefficients
turbines by NREL FAST and 9.1 RPM by GH Bladed. These results in Cl, Cd, Cn and Ct are computed evaluating forces on a section of the
significant decreases in standard deviation for the power, by a foil and using the same velocity used to compute the AOA. The
factor of 9 for NREL FAST and 6 for GH Bladed, see cases 2.1a and computations are conducted in the airfoil system of reference, so
2.1b in Table 4. The turbine can then be operated with more stable transformations are applied to project velocities from the earth
power output. Increasing the rotor speed at constant wind in- reference system, where velocities are computed by CFD, to the
creases tip speed ratio and thus usually helps to improve efficiency. blade reference system. These transformations account for yaw, tilt,
However, the torque decreases slightly for the case with controller, precone, and rotor angle.
by 3% for NREL FAST and 1.4% for GH Bladed. The thrust increases Fig. 14 shows time histories of the AOA, a and a0 for blade 1 at 5
about 2.5% for NREL FAST and 0.8% for GH Bladed. In this case of selected radial positions for wind speed of 8 m/s. Notice that at r/
moderate wind speeds, higher rotor speed results in larger thrust R ¼ 0.16 the method by Shen et al. [45] becomes unstable due to
and smaller torque for the flexible turbine. separation, resulting in noise in the prediction of the AOA. The
Fig. 11 and Table 5 show turbine thrust and torque for rigid and presence of the tower is clear when the blade is on the bottom
flexible blades and tower. Comparing with its rigid turbine coun- (integer number of rotations) by an increase in axial induction
terpart, the CFD approach predicts an increase of 3.5% for average factor and decrease in angle of attack. Since the tilt moves the blade
thrust and 0.6% decrease for torque, in close agreement with cases downstream when moving from bottom to top, the incoming wind
2.1b and 3.1 for the OC3 results. For the flexible turbine the mean velocity in the airfoil system of reference decreases as the blade
magnitude of thrust predicted by CFD is 6.3% higher than results moves from bottom to top and increases when moving from top to
from GH Bladed, while NREL FAST predicts 8% more than GH bottom. At the same time, when the blade is at the top it is
Bladed. The average torque is 0.5% lower than GH Bladed, and NREL immersed deeper into the wake, resulting in lower axial velocities.
FAST is 4.5% higher. Due to the lack of a controller, the standard As a consequence, for all sections with r > 0.5R, the AOA has a
deviation predicted for the loads by CFD are much larger than those trough in uniform wind approximately when the blade is at the top
in OC3 participants. Since the OC3 results deviated from the and a high when it is on the bottom. The axial induction factor is
nominal rotor speed of 9 RPM due to the use of the controller, as highest when the blade is at the top and a minimum for the blade
discussed in the previous paragraph, the differences between OC3 on the bottom, approximately amax ¼ 0.4 and amin ¼ 0.33 at r/
participants and the CFD results could be higher or lower than R ¼ 0.93. The AOA is below the stall angle of attack (shown as insets
those reported in this study. in Fig. 14) for all sections except r/R ¼ 0.16 at the root, 0.28 < a < 0.4
Loads for the individual blades for the rigid and flexible cases are and a0 < 0.1, indicating operation of the turbine close to design
shown in Fig. 12. Neglecting the effect of the tower, which causes a point.
sudden drop in thrust and torque when the blade is at the bottom, For uniform wind, a flexible blade results in slightly higher
there is a minimum in loads when the blade is approximately at the AOAs, a lower induction factor for sections closer to the shaft and a
top, and a maximum when the blade is approximately at the bot- higher induction factor near the tip. The higher axial induction
tom. This behavior can be explained in terms of changes in the factor near the tip of the blade is expected as the blades deform and
angle of attack due to flow and geometry, as discussed below in gets deeper into the wake. Due to deflections, the blade twist de-
reference to Fig. 14. creases a small amount close to the root to about 1 in average at
Classical BEM characterize the airfoil sections in terms of pa- the tip, resulting in larger AOAs for flexible blades.
rameters describing local flow and forces. These include the The trends seen in AOA and a determine the blade load behavior
sectional angle of attack (AOA), axial induction factor (a) and shown in Fig. 12. The blade thrust is the integration of the differ-
tangential induction factor (a0 ), lift (Cl) and drag (Cd) coefficients, ential normal force Fn ¼ Cn  (1/2rVrel 2
c), where Cn is the normal
normal (Cn) and tangential (Ct) coefficients. CFD results can be force coefficient, Vrel is the relative velocity seen by the airfoil and c
inspected to obtain these airfoil parameters, though the definition is the chord length. For AOAs below stall, Cn is approximately
of some of them makes evaluation somewhat ambiguous. Fig. 13 proportional to AOA, while ignoring tilt and precone
shows the streamlines at 4 different spanwise sections of blade 1 2
Vrel ¼ (V0(1  a))2 þ (ur(1 þ a0 ))2. Note that this approximation is
colored by dimensionless axial velocity, exhibiting streamlined only used here for the sake of simplicity. Examining Figs. 12 and 14,
flow on all sections except near the root. Flow velocities were the increase in AOA for a flexible turbine results in larger Cn than for
Y. Li et al. / Renewable Energy 76 (2015) 338e361 353

Fig. 13. Instantaneous flow field contoured by non-dimensional axial velocity in blade system at wind speed 8 m/s.

Fig. 14. Time history of angle of attack (left), axial induction factor a (center) and tangential induction factor a0 (right) at selected radial positions for blade 1 with wind speed 8 m/s.
354 Y. Li et al. / Renewable Energy 76 (2015) 338e361

Fig. 15. Average pressure coefficient and standard deviation at selected radial positions for wind speed 8 m/s at uniform wind (a) and log-law wind (b) profiles (solid: rigid turbine;
dashed: flexible turbine).

a rigid turbine, with consequent higher thrust. In addition, AOA and Fig. 12. Ct exhibits less differences between rigid and flexible re-
(1  a)2 determine the trend of Fn and thus the blade thrust, sults, consistent with the torque behavior shown in Fig. 12.
showing periodic oscillations with primary trough when blade is at Fig. 17 shows the lift coefficient Cl and AOA over two rotor rev-
the top and secondary trough when it is close to the tower, peaking olutions for the rigid (Fig. 17(a)) and flexible (Fig. 17(b)) turbines.
approximately 0.1 rotation before and after passing the tower. The steady-state angle of attack-lift coefficient curve is shown in
Fig. 15 shows the average and standard deviation of the pressure Fig. 17(c). Note that the zero azimuthal angle in Fig. 17(c) corre-
coefficient Cp for the rigid and flexible turbines at 5 different radial sponds to the blade pointing downwards (tower passage), while
sections. The section at r/R ¼ 0.16 is very close to the root and in the 180 corresponds to the blade atop. As shown in Fig. 17(a) and (b),
transition from cylindrical to airfoil section. From 60% chord to the both rigid and flexible turbines display a small delay between Cl
trailing edge, Cp at the pressure side is slightly smaller than at the and AOA. Time changes in AOA and aerodynamic loads are caused
suction side, indicating that in that portion of the section the lift by tilt angle effects, as discussed before, resulting in changes of the
coefficient is negative. At the same time, this section also shows AOA of amplitude determined by the tilt angle and a frequency
large lift coefficient from 0 to 60% chord. However, the strong determined by the rotation frequency. In addition, the presence of
adverse pressure gradient after 20% chord is an indicator of possible the tower causes additional transient in AOA and loads. Similar
flow transition and separation. This is consistent with the high AOA dynamic effects exhibiting time lag and hysteresis were observed in
shown for that section in Fig. 14. Standard deviations for Cp at this an experiment conducted by Fuglsang et al. [46] in which the dy-
section are large, further confirming the unsteadiness and separa- namic AOA is controlled by a pitching motion mechanism with
tion. All other sections show standard behavior of attached flow, imposed amplitude and frequency, achieving similar effects as
with small standard deviation and moderate pressure gradients on caused by tilt. Fig. 17(c) shows primarily a counter clockwise path in
the suction side. Note that pressure coefficients for the elastic and all sections except at r/R ¼ 0.16, indicating a smooth flow and the
rigid turbines have almost identical trends. absence of dynamic stall. When the blade rotates from the top and
Fig. 16 shows radial distribution of average and standard devi- approaches the tower between 270 and 0 , a secondary counter
ation of the normal and tangential coefficients Cn and Ct for the clockwise path is formed as the blade passes the tower. Paths for
rigid and flexible turbines. Since the structure is very rigid close to the rigid and flexible turbines stay close at inboard sections, but
the root, there is no significant difference between rigid and flexible then gradually depart from each other, as can be seen in the
results, but differences are apparent for higher r/R for Cn. Cn is partially intersected loops at r/R ¼ 0.55 and fully separated at r/
essentially flat with mean magnitude around 1.0 for 0.6 < r/R < 0.8, R ¼ 0.93.
where most of the torque is produced, while the flexible turbine The drag coefficient Cd, not pictured, shows similar trends as the
gradually increases from 1.0 at r/R ¼ 0.6 to 1.1 at r/R ¼ 0.8. The lift coefficient, with the particularity that Cd in the section r/R ¼ 0.16
losses at the tip are clear for both rigid and flexible turbine, but the is about 10 times higher than at other sections, due to the much
flexible turbine shows larger Cn. This explains the large difference thicker geometry of the airfoil in the root/airfoil transition region,
for the blade thrust between rigid and flexible blade as shown in and also exhibiting significant fluctuations as observed for the lift.

Fig. 16. Radial distribution of Cn and Ct for wind speed 8 m/s.


Y. Li et al. / Renewable Energy 76 (2015) 338e361 355

Fig. 17. AOA and Cl for wind speed 8 m/s.

Fig. 18 and Table 6 compare the predicted blade 1 tip deflections 3.011 m, respectively. The maximum deflection is 3.692 m for CFD,
from CFD and by OC3 participants. The deflections are respect to 3.37 m for NREL FAST and 3.149 m for GH Bladed.
the coned coordinate system that rotates with the rigid rotor. For the In-Plane (IP) deflections, all three methods predict
Positive out-of-plane (OoP) deflection points downwind in the essentially sinusoidal patterns, determined mainly by centripetal
coned coordinate system, positive in-plane (IP) deflection points force, blade rotation and tilt effects, but dominated by gravity. Due
from leading edge to trailing edge, opposite to the blade rotational to the CM offset to the blade pitch axis and the twist angle, cen-
direction. Due to the rotational and tilt effects, all blade deflections tripetal forces cause a negative deflection towards leading edge,
exhibit cyclic oscillations. while aerodynamic forces from the wind push the curve to negative
OoP deflection is important since it is related to the structural deflections, leading to asymmetric oscillations with smaller
strength and fatigue of the blade, and to tower clearance issues. deflection towards trailing edge and larger deflection towards
Since thrust is the largest contributor to the OoP tip deflection, the leading edge. CFD predicts an average 0.345 m for IP deflections,
predicted deflection follows similar trends as those observed and while NREL FAST and GH Bladed show an average value of 0.318 m
discussed in Fig. 12 for blade thrust. The primary trough occurs and 0.304 m, respectively. All three methods also show similar
approximately when blade is at the top with the least influence deflection amplitudes.
from gravity, and the secondary trough is induced by the tower
shadow with mean magnitude of about 3.5 m. If the tower was not 4.4. Effect of wind shear
present, the deflection would keep increasing to a maximum value
with the blade on the bottom where the gravity effect is maximum Realistic wind profiles account for the atmospheric boundary
due to the tilt and precone angle. As a consequence, the tower layer shear and turbulence. In this paper, wind shear is modeled
shadow contributes to increase the blade/tower clearance. As seen with the log-law wind profile as inlet boundary condition,
in Table 6, CFD shows OoP mean deflection of 3.592 m, while NREL u(z) ¼ u(zref)ln(z/z0)/ln(zref/z0), where zref is the hub height and z0 is
FAST and GH Bladed predict averaged deflections of 3.244 m and the surface roughness. The surface roughness is taken as the value
356 Y. Li et al. / Renewable Energy 76 (2015) 338e361

Fig. 18. Deflections for blade 1 tip for wind speed 8 m/s.

for a smooth sea condition, z0 ¼ 104 m. A rougher surface will turbulence. The anisotropic model parameters used in Section 4.1
result in larger z0 and the larger wind shear. With the non- step 2 were applied to all simulations in turbulent wind. Elec-
dimensional axial wind speed at hub height as 1, the wind speed tronic Annex I shows a video comparing the cases with and without
experienced by a blade of 63 m long at the tip is 0.912 at lowest turbulence for a flexible turbine.
position and 1.039 at the top (ignoring tilt), a 14% difference. Such Supplementary video related to this article can be found at
difference exposes the airfoil to changing inflow conditions within http://dx.doi.org/10.1016/j.renene.2014.11.014.
a blade rotation, which subsequently leads to significant different Note that no time history results were reported by OC3 partic-
blade load behavior than with uniform wind profile. ipants for the wind turbulence cases, but statistics and results in
Fig. 11 compares the time history of turbine thrust and torque frequency domain were reported, allowing for some comparison.
for uniform and log-law wind and Table 5 shows quantitative Also note that the OC3 cases have a pitch controller which signifi-
values. The general trends are similar to those with uniform wind, cantly changes the behavior of the turbine, making comparisons
but the mean thrust is about 1% lower and the torque about 2% with fixed pitch CFD only qualitative.
lower for both rigid and flexible turbines. The individual blade Fig. 19 shows time histories of thrust and torque with wind
behavior, shown in Fig. 12, exhibits a significant increase in thrust turbulence, including a case for rigid turbine without wind turbu-
and torque when the blade is at the top and a decrease when the lence for reference. The corresponding statistics are shown in
blade is at the bottom, as expected. A logarithmic wind velocity Table 7, including OC3 results [40]. The rigid turbine with wind
distribution immediately results in higher AOAs when the blade is turbulence shows small increases of 0.2% in mean thrust and 1.4% in
at the top and the incoming wind speed is higher, with a dramatic mean torque respect to the case without turbulence, while the
decrease when the blade is at the bottom adding the effect of lower elastic turbine sees an increase of 4.7% for thrust and decrease of
wind speed with the tower shadow effect, see Fig. 14. Conversely, 4.9% for torque under wind turbulence, consistent with the trends
the axial induction factor tends to be flatter than for uniform winds.
Due to the blade loads affected by wind shear, blade OoP de-
flections under log-law wind follow closely the individual blade
Table 6
loads shown in Fig. 14, see Fig. 18 for time history of deflections and
Blade 1 tip deflections for wind speed 8 m/s.
Table 6 for mean and standard deviation values. IP deflections are
less affected by the vertical wind distribution due to the large edge- Participant CFD NREL FAST GH Bladed

wise stiffness of the blade, but the deflection from trailing edge to Case Uniform wind Log-law 3.1 3.1
leading edge (negative IP) decreases by a small amount. Out-of-plane Mean 3.592 3.559 3.244 3.011
deflection [m] Min 3.510 3.309 3.109 2.857
Max 3.692 3.709 3.370 3.149
4.5. Effect of wind turbulence s 0.059 0.107 0.090 0.098
In-plane Mean 0.345 0.340 0.318 0.304
The Mann wind turbulence model described in Section 2.4 was deflection [m] Min 0.703 0.702 0.795 0.749
applied to both rigid and flexible turbines at the rated condition of Max 0.001 0.007 0.164 0.139
s 0.246 0.249 0.334 0.311
11.4 m/s with uniform wind to investigate effect of wind
Y. Li et al. / Renewable Energy 76 (2015) 338e361 357

Fig. 19. Thrust and torque for wind speed 11.4 m/s.

seen in OC3 results. Large fluctuations exist due to the wind only to turbulence and therefore relatively weak. The zero-
turbulence, with approximately 7 times larger standard deviation frequency component of the PSD is thus related to the average of
for both thrust and torque compared to the no wind turbulence the variable (thrust or torque). From Fig. 20, though deviation exists
case. between CFD and OC3 participants due to the controller, the results
Power spectrum densities (PSD) for thrust and torque are shown for both rigid and flexible turbine are very similar in the frequency
in Fig. 20. CFD data from the 5th to 19th rotation was used to domain. All simulation results show a primary peak at around
perform the one-sided PSD estimates using FFT. No periodicity was 0.6 Hz, corresponding to the blade passage frequency, and then
enforced explicitly, but deviations from periodic behavior are due followed by peaks at higher harmonics of the blade passage fre-
quency (1.2 Hz, 1.8 Hz). Since the integral of the power spectrum
over all frequencies is the corresponding statistical variance, OC3
results in general show slightly higher energy content for thrust
Table 7 and noticeable lower energy content for torque. For operation at
Thrust and torque for wind speed 11.4 m/s, including OC3 results for FAST and GH the rated condition (11.4 m/s), a control system was used by OC3
Bladed.
participants while CFD simulations use the rated rotor speed of 12.1
Participant CFD NREL FAST GH Bladed RPM. The controller uses generator torque control for wind speeds
Case No Wind 2.2 2.2 below 11.4 m/s, and blade pitch control to maintain the rated power
turbulence, turbulence, for wind speed beyond the rated condition, resulting in the use of
rigid rigid the generator torque controller for the uniform wind case and both
Thrust [kN] Mean 758.7 760.0 606.4 576.2 generator torque and blade pitch controllers for the wind turbu-
Min 737.3 640.0 365.0 273.9 lence case. The consequence is that the power control reduces
Max 768.8 858.0 835.8 864.3 fluctuations in torque for the OC3 results. Additionally, OC3 results
s 7.23 48.88 82.83 89.40
Shaft torque Mean 4267.34 4327.05 4031.00 3895.73
show smoother curves due to the larger frequency window
[kN m] Min 4003.72 2890.96 3015.00 2606.10 (0.0333 Hz) than CFD (0.0144 Hz).
Max 4385.58 5706.80 4410.00 4531.79 Fig. 21 compares blade tip deflections for 11.4 m/s with and
s 88.59 619.86 263.50 430.10 without wind turbulence. The OoP deflection increased consider-
ably respect to the 8 m/s case shown in Fig. 18, where the average
Case Wind turbulence, flexible 3.2 3.2
OoP deflection was approximately 3.6 m, increasing now to 6.3 m.
Thrust [kN] Mean 795.8 633.4 588.4 The turbulence-induced fluctuations are considerable, reaching a
Min 694.1 409.7 358.1
maximum of almost 8 m, still far from the blade-tower design
Max 869.9 855.1 855.7
s 44.40 84.51 79.81 clearance of 13.2 m. Turbulence effects on IP deflection are much
Shaft torque Mean 4113.9 4010.0 3873.6 smaller, but still clearly noticeable. The power spectrum density of
[kN m] Min 2983.4 2896.0 2583.1 blade tip OoP and IP deflections are shown in Fig. 22. Peaks at low
Max 5067.5 4698.0 4625.7 frequency compare well, but at higher frequencies CFD losses en-
s 519.06 301.90 444.62
ergy faster. Unfortunately a fair comparison is not possible since the
358 Y. Li et al. / Renewable Energy 76 (2015) 338e361

Fig. 20. Power spectrum density of thrust and torque (wind speed ¼ 11.4 m/s).

Fig. 21. Blade tip deflections for wind speed 11.4 m/s.

Fig. 22. Power spectrum density of blade tip deflections for wind speed 11.4 m/s.
Y. Li et al. / Renewable Energy 76 (2015) 338e361 359

Fig. 23. Vortical structures represented by Q ¼ 1 for cases with and without wind turbulence, for mean hub height wind speed 11.4 m/s.

Fig. 24. Axial velocities at different axial positions and at hub height (z ¼ 90 m, z/R ¼ 1.428, top) and at top tip height (z ¼ 153 m, z/R ¼ 2.428, bottom) for wind speed 11.4 m/s.
360 Y. Li et al. / Renewable Energy 76 (2015) 338e361

use of a controller by OC3 introduces frequencies not present in was examined for the wake flow to analyze the influence of wind
CFD. turbulence on wake diffusion, finding that explicit turbulence
addition results in considerably increased wake diffusion.
4.6. Flow field Future work will focus on the development of a gearbox model
that enables investigation of interaction between aerodynamic
Vortical structures for the cases with and without turbulence loads and gearbox reaction including shaft-gearbox-generator dy-
are shown in Fig. 23 for a wind speed of 11.4 m/s. The vortical namics, allowing prediction of gear-level loads resulting from tur-
structures rendered as isosurfaces of Q ¼ 1, colored with axial ve- bulent wind fluctuations. Also, studies of the offshore wind turbine
locity. The case without wind turbulence has strong tip and hub under incoming waves with the mooring lines model recently
vortices, as well as vortex detachment from the tower. Notice the implemented by Ref. [47] will be performed.
tip-type vortices forming at the gap in the grid between the hub
and the blade root. The tip vortices cannot be resolved beyond x/
Acknowledgments
R ¼ 1, the region covered by the finer refinement grid. In the case
with turbulence vortical structures are present in the system as
This research was supported by IAWIND (Grant IOEI 188570301)
part of external atmospheric turbulence. These turbulent structures
under project “Tool Development for Direct Simulation of Interac-
interact with the turbine, in particular with the tip vortices which
tion between Aerodynamic and Gearbox Loads,” and the National
become unstable and breakdown closer to the rotor. External wind
Science Foundation under grant 1066627, with Dr. Ram Gupta as
turbulence also diffuses the wake faster, as can be seen comparing
manager of the Energy for Sustainability Program. Computations
the cross sections at y/R ¼ 0.8 for the cases with and without wind
were performed in Helium, the high performance computing
turbulence.
resource at the University of Iowa.
Cross sections of the wake are shown in Fig. 24. Axial velocity on
lateral traverses at hub and top tip height for x/R ¼ 1, 1, 2, 3, 4 are
plotted. The CFD results show that the flexibility of the turbine has References
little effect on the wake, an important point since it suggests that
[1] Spinato F, Tavner PJ, Van Bussel GJW, Koutoulakos E. Reliability of wind tur-
wake studies can be performed with reasonable approximation bine subassemblies. IET Renew Power Gener 2009;3:387e401.
using rigid turbines. The wind turbulence, on the other hand, has a [2] Sørensen JN, Shen W, Munduate X. Analysis of wake states by a full-field
considerable effect in the wake by adding, as expected, an effective actuator disc model. Wind Energy 1998;1:73e88.
[3] Sørensen JN, Mikkelsen R, Troldborg N. Simulation and modelling of turbu-
eddy viscosity that diffuses the wake faster. This effect highlights
lence in wind farms. In: Proceedings EWEC 2007; 2007.
the importance of proper estimation of the atmospheric and local [4] Sørensen JN, Shen W. Numerical modeling of wind turbine wakes. J Fluids Eng
turbulence to properly predict the wakes in a wind farm. 2002;124:393.
[5] Shen W, Zhang J, Sørensen JN. The actuator surface model: a new Naviere-
Stokes based model for rotor computations. J Sol Energy Eng 2009:131.
5. Conclusions and future work [6] Potsdam MA, Mavriplis DJ. Unstructured mesh CFD aerodynamic analysis of
the NREL phase VI rotor. In: 47th AIAA aerospace sciences meeting; 2009.
A high fidelity approach for wind turbine aero-elastic simula- [7] Duque EP, Burklund MD, Johnson W. NaviereStokes and comprehensive
analysis performance predictions of the NREL phase VI experiment. J Sol En-
tions including explicit representation of the atmospheric wind ergy Eng 2003;125:457e67.
turbulence has been presented. The approach uses a dynamic [8] Bazilevs Y, Hsu M-C, Akkerman I, Wright S, Takizawa K, Henicke B, et al. 3D
overset CFD code to compute the aerodynamics coupled with an simulation of wind turbine rotors at full scale. Part I: geometry modeling and
aerodynamics. Int J Numer Methods Fluids 2011;65:207e35.
MBD code to predict the motion responses to the aerodynamic [9] Hansen M, Sørensen JN, Voutsinas S, Sørensen N, Madsen HA. State of the art
loads. The IEC 61400-1 ed. 3 recommended Mann wind turbulence in wind turbine aerodynamics and aeroelasticity. Prog Aerosp Sci 2006;42:
model was implemented into the CFD code as boundary and initial 285e330.
[10] Bauchau OA. Computational schemes for flexible, nonlinear multi-body sys-
conditions. The wind turbulence model was validated by
tems. Multibody Syst Dyn 1998;2:169e225.
comparing the generated stationary wind turbulence field with the [11] Bauchau OA, Bottasso CL, Nikishkov YG. Modeling rotorcraft dynamics with
theoretical one-point spectrum for the three components of the finite element multibody procedures. Math Comput Model 2001;33:1113e37.
[12] Biedron RT, Lee-Rausch EM. Rotor airloads prediction using unstructured
velocity fluctuations, and by comparing the expected statistics from
meshes and loose CFD/CSD coupling. In: 26th AIAA applied aerodynamics
the CFD simulated wind turbulent field with the explicit wind conference, Honolulu, Hawaii, United States; 2008.
turbulence inlet boundary from Mann model. Extensive simula- [13] Corson D, Griffith DT, Ashwill T, Farzin S. Investigating aeroelastic perfor-
tions based on the proposed coupled approach were conducted mance of multi-megawatt wind turbine rotors using CFD. In: 53rd AIAA/
ASME/ASCE/AHS/ASC structures, structural dynamics and materials confer-
with the conceptual NREL 5-MW offshore wind turbine in an ence 20th AIAA/ASME/AHS adaptive structures conference 14th AIAA, Hon-
increasing level of complexity, so as to validate and analyze the olulu, Hawaii, United States; 2012.
aerodynamic predictions, elasticity, wind shear and atmospheric [14] Bazilevs Y, Hsu M-C, Kiendl J, Wüchner R, Bletzinger K-U. 3D simulation of
wind turbine rotors at full scale. Part II: fluidestructure interaction modeling
wind turbulence. Results were compared with the publicly avail- with composite blades. Int J Numer Methods Fluids 2011;65:236e53.
able simulations results from OC3 participants, showing good [15] Hsu M-C, Bazilevs Y. Fluidestructure interaction modeling of wind turbines:
agreement for the aerodynamic loads and blade tip deflections in simulating the full machine. Comput Mech 2012;50:821e33.
[16] Smirnov A, Shi S, Celik I. Random flow generation technique for large eddy
time and frequency domains. simulations and particle-dynamics modeling. Trans ASME-I-J Fluids Eng
The study shows that the turbine is well designed with mostly 2001;123:359e71.
attached flow along the blade under normal operational conditions, [17] Bechmann A. Large-eddy simulation of atmospheric flow over complex
terrain [Ph. D diss.], Risø-PhD-28(EN). Risø National Laboratory, Technical
showing angle of attack, axial induction factor, lift coefficient and University of Denmark; 2006.
airfoil pressure distributions within high-performance range. The [18] Veers PS. Three-dimensional wind simulation. No. SAND-88-0152C, CONF-
analysis with a log-law wind profile inflow shows its strong influ- 890102-9. Albuquerque, New Mexico, United States: Sandia National Labo-
ratories; 1988.
ence on the blade thrust and torque and deflections, pointing to the
[19] Mann J. The spatial structure of neutral atmospheric surface-layer turbulence.
need of considering realistic wind profiles when designing large- J Fluid Mech 1994;273:141e68.
scale wind turbines. A similar conclusion can be drawn from the [20] Mann J. Wind field simulation. Probab Eng Mech 1998;13:269e82.
[21] Troldborg N, Zahle F, Rethore P-E, Sørensen NN. Comparison of wind turbine
results including atmospheric turbulence. The CFD results also
wake properties in non-sheared inflow predicted by different computational
suggest that wake studies can be performed with reasonable ac- fluid dynamics rotor models. Wind Energy 2014. http://dx.doi.org/10.1002/
curacy using rigid turbines. Wind turbulence/turbine interaction we.1757.
Y. Li et al. / Renewable Energy 76 (2015) 338e361 361

€mer E. CFD studies on wind turbines in


[22] Schulz C, Klein L, Weihing P, Lutz T, Kra [35] Tang HS, Casey Jones S, Sotiropoulos F. An overset-grid method for 3D un-
complex terrain under atmospheric inflow conditions. J Phys Conf Ser steady incompressible flows. J Comput Phys 2003;191:567e600.
2014;524:012134. [36] Pope SB. Turbulent flows. Cambridge University Press; 2000.
[23] International Electrotechnical Committee. IEC 61400-1: wind turbines part 1: [37] Frigo M, Johnson S. The design and implementation of FFTW3. Proc IEEE
design requirements. International Electrotechnical Commission; 2005. 2005;93:216e31.
[24] Prescott WC. Numerical integration technique for multi-body dynamic system [38] Jonkman JM, Butterfield S, Musial W, Scott G. Definition of a 5-MW reference
software. In: Symposium on international automotive technology, Pune; wind turbine for offshore system development. NREL/TP-500-38060. National
1999. Renewable Energy Laboratory; 2009.
[25] LMS International. Virtual.Lab online help manual. 2013. [39] Jonkman JM. Definition of the floating system for phase IV of OC3. NREL/TP-
[26] Haug EJ. Computer aided kinematics and dynamics of mechanical systems. 500-47535. National Renewable Energy Laboratory; 2010.
Boston: Allyn and Bacon; 1989. [40] Jonkman JM, Musial W. Offshore Code Comparison Collaboration (OC3) for IEA
[27] Li Y, Paik K-J, Xing T, Carrica PM. Dynamic overset CFD simulations of wind task 23 offshore wind technology and deployment. NREL/TP-5000-48191.
turbine aerodynamics. Renew Energy 2012;37:285e98. National Renewable Energy Laboratory 2010
[28] Huang J, Carrica PM, Stern F. Semi-coupled air/water immersed boundary [41] Laino DJ, Hansen AC, Minnema JE. Validation of the AeroDyn subroutines
approach for curvilinear dynamic overset grids with application to ship hy- using NREL unsteady aerodynamics experiment data. Wind Energy 2002;5:
drodynamics. Int J Numer Methods Fluids 2008;58:591e624. 227e44.
[29] Carrica PM, Wilson RV, Stern F. An unsteady single-phase level set method for [42] Jonkman JM, Buhl Jr ML. FAST user's guide. NREL/EL-500-29798. National
viscous free surface flows. Int J Numer Methods Fluids 2007;53:229e56. Renewable Energy Laboratory; 2005.
[30] Carrica PM, Wilson RV, Noack RW, Stern F. Ship motions using single-phase [43] Marple Jr SL. Digital spectral analysis with applications. Englewood Cliffs, NJ:
level set with dynamic overset grids. Comput Fluids 2007;36:1415e33. Prentice-Hall, Inc.; 1987. 512 p.
[31] Noack RW. SUGGAR: a general capability for moving body overset grid as- [44] Larsen TJ. Turbulence for the IEA annex 30 OC4 project. Risø-I-3206(EN). Risø
sembly. In: 17th AIAA computational fluid dynamics conference, Toronto, National Laboratory; 2013.
Ontario, Canada; 2005. [45] Shen WZ, Hansen MOL, Sørensen JN. Determination of the angle of attack on
[32] Carrica PM, Huang J, Noack R, Kaushik D, Smith B, Stern F. Large-scale DES rotor blades. Wind Energy 2009;12:91e8.
computations of the forward speed diffraction and pitch and heave problems [46] Fuglsang P, Antoniou I, Sørensen N, Madsen H. Validation of a wind tunnel
for a surface combatant. Comput Fluids 2010;39:1095e111. testing facility for blade surface pressure measurements. Risø-R-981(EN). Risø
[33] Gritskevich MS, Garbaruk AV, Schütze J, Menter FR. Development of DDES and National Laboratory; 1998.
IDDES formulations for the keu shear stress transport model. Flow, Turbul [47] Quallen S, Xing T, Carrica PM, Li Y, Xu J. CFD simulation of a floating offshore
Combust 2011;88:431e49. wind turbine system using a quasi-static crowfoot mooring-line model. In:
[34] Menter FR. Two-equation eddy-viscosity turbulence models for engineering 23rd International ocean and polar engineering conference (ISOPE-2013),
applications. AIAA J 1994;32:1598e605. Anchorage, AK, United States; 2013.

You might also like