Advanced CFD Methods For Wind Turbine Analysis
Advanced CFD Methods For Wind Turbine Analysis
A Dissertation
Presented to
The Academic Faculty
by
C. Eric Lynch
In Partial Fulfillment
of the Requirements for the Degree
Doctor of Philosophy in the
School of Aerospace Engineering
Approved by:
iii
ACKNOWLEDGEMENTS
First and foremost, I would like to thank my advisor, Dr. Marilyn Smith for her guidance
through my years in graduate school. Thanks also go to the other members of my committee
This work was supported in part by the National Science Foundation, Project 0731034,
“Advancing Wind Turbine Analysis and Design for Sustainable Energy”. I would like to
thank NSF and the NSF Program Officers, Dr. Trung Van Nguyen and Dr. Greg Rorrer,
for their support in this endeavor. Any opinions, findings, and conclusions or recommenda-
tions expressed in this material do not necessarily reflect the views of the National Science
Foundation. Computational support for this research was supported in part by the National
were hosted by NCSA, LONI, and TACC. Sincere thanks go Dr. Christopher Stone for his
I would like to thank Bob Biedron and Beth Lee-Rausch at NASA Langley for providing
the HART grids used in this work as well as the data from their HART simulations for
comparison. Thanks also go to Scott Schreck at NREL for providing the NREL UAE Phase
VI datasets.
I would like to thank the members of the Fun3d development team for allowing me
access to the Fun3d SVN repository so I could share the HRLES turbulence model and
kd-tree nearest neighbor search with other users. I also thank Ralph Noack of the Penn
State ARL for all his assistance using Suggar++ (and putting up with my bug reports).
both past and present, for all their help over the years. In particular, I would like to thank
Jennifer Abras and Nic Reveles for numerous helpful discussions about CFD/CSD coupling.
Thanks also go to Rajiv Shenoy for his help with grid generation and Jean de Montaudouin
iv
TABLE OF CONTENTS
DEDICATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii
ACKNOWLEDGEMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv
SUMMARY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi
I MOTIVATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
II BACKGROUND . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.1 Actuator Disc and Blade Techniques . . . . . . . . . . . . . . . . . . . . . 5
2.2 Hybrid RANS/LES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.3 Full Blade and Rotor CFD . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.4 CFD-CSD Coupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.4.1 Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.4.2 Development of Early Coupling Methods . . . . . . . . . . . . . . 16
2.4.3 Current State of the Art . . . . . . . . . . . . . . . . . . . . . . . 18
V HYBRID RANS/LES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
5.1 Governing Equations of Motion . . . . . . . . . . . . . . . . . . . . . . . . 42
5.2 Hybrid RANS-LES Model . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.3 Unstructured Implementation . . . . . . . . . . . . . . . . . . . . . . . . . 47
v
5.4 Method Verification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
5.4.1 Initial Evaluation: Circular Cylinder . . . . . . . . . . . . . . . . . 48
5.4.2 Unstructured Mesh Studies . . . . . . . . . . . . . . . . . . . . . 51
IX CONCLUSIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
9.1 Recommendations for Future Work . . . . . . . . . . . . . . . . . . . . . 104
REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
VITA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
vi
LIST OF TABLES
vii
LIST OF FIGURES
viii
14 Instantaneous contours of the blending function of Eq. 17 for the 2-D circular
cylinder. Lighter areas are RANS-dominated, and darker areas are LES-
dominated. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
15 Computational grid for the 3-D cylinder. . . . . . . . . . . . . . . . . . . . . 52
16 Iso-surfaces of Q criterion about a 3-D circular cylinder with varying grid
resolution. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
17 Mean pressure coefficient along the centerline for the 3-D circular cylinder
on several grids with both RANS and HR-LES turbulence models. LES data
is from Reference [40], and experimental data is from Reference [65]. . . . . 55
18 Correlation of computed mean pressure coefficient with measured value. . . 55
19 Iso-vorticity contours around a 3-D cylinder, computed using the HR-LES
model on grids with different cell types in the boundary layer. . . . . . . . . 56
20 u/U∞ at seven locations in the wake of a cylinder at Re = 3900 using a
spanwise extruded grid. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
21 Spanwise slices of the original mesh and the 3-D mesh used to evaluate the
effect of cell stretching. Red lines indicate wake traverse locations. . . . . . 59
22 u/U∞ at seven locations in the wake of a cylinder at Re = 3900 using a fully
3-D grid. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
23 Grids used for DU97 simulations with wind tunnel walls. These 2-D grids
were extruded in the spanwise direction. . . . . . . . . . . . . . . . . . . . . 63
24 Isosurfaces of Qc/â∞ = 1.012 around the DU97 airfoil in the wind tunnel at
α = 10◦ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
25 Vorticity magnitude at midspan for the DU97 airfoil in wind tunnel at α = 10◦ 65
26 Normalized frequency spectrum of the lift history for the DU97 case. The
dual peaks in (b) indicate a secondary shedding mode. . . . . . . . . . . . . 66
27 Structured overset on DU97 airfoil. Figure and grids taken from Ref. [104]. 67
28 Q criterion iso-surfaces around the overset DU97 airfoil with free boundaries
(grid 3). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
29 Hub of the HART Dymore model. . . . . . . . . . . . . . . . . . . . . . . . 71
30 Convergence of the 1st torsion frequency with the number of finite elements
in the blade beam. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
31 Fan plot for the HART Dymore model. . . . . . . . . . . . . . . . . . . . . 73
32 First six mode shapes for the HART Dymore model. . . . . . . . . . . . . 74
33 13.6 million node composite mesh used for HART-II simulations. . . . . . . 75
34 Convergence of blade pitch controls with loose coupling iterations. . . . . . 77
ix
35 Convergence of airloads at 87% span for the loosely coupled trim phase. For
measured pitching moment, the mean value is -0.00258; for iteration 5, the
mean pitching moment is -0.00513. . . . . . . . . . . . . . . . . . . . . . . . 79
36 Comparison of structural moments computed using Fun3d/Dymore and
Fun3d/Camrad in Reference [16]. Mean values removed in second column
are listed in Table 6. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
37 Comparison of airloads at 87% span computed using Fun3d/Dymore and
Fun3d/Camrad in Reference [16]. Mean pitching moments are -0.00258, -
0.00424, and -0.00513 for measured, Fun3d/Camrad, and Fun3d/Dymore,
respectively. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
38 Tip deflections predicted in tight coupling and the final iteration of loose
coupling. Bars on the measured data indicate blade-to-blade variation. . . . 82
39 Sectional normal force and pitching moment 87% span, comparing loose and
tight coupling. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
40 Q = 0.0075 iso-surfaces. Q is non-dimensionalized by speed of sound and
reference length, 1 m. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
41 NREL Phase VI turbine mounted in 80 × 120’ tunnel. . . . . . . . . . . . . 86
42 The final composite mesh for the NREL Phase VI turbine. . . . . . . . . . . 88
43 Mean sectional pressure coefficient at 10 m/s. Bars indicate standard devi-
ation of pressure at a particular tap, not error. RANS and HR-LES results
are coincident. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
44 Q-criterion iso-surface (Q = 0.1) at 10 m/s, predicted using HR-LES. . . . . 92
45 Sectional pressure coefficient at 15 m/s, 0 yaw. . . . . . . . . . . . . . . . . 93
46 Q-criterion iso-surface (Q = 0.1) at 15 m/s, predicted using RANS. . . . . . 94
47 Q-criterion iso-surface at 15 m/s, predicted using HR-LES. . . . . . . . . . 95
48 Sectional pressure coefficient at 15 m/s, 30◦ yaw. . . . . . . . . . . . . . . . 96
49 Integrated hub loads (pressure component only) vs. azimuth at 15 m/s, 30◦
yaw. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
50 Sectional normal force coefficient, normalized by tip dynamic pressure, vs.
azimuth at 15 m/s, 30◦ yaw. . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
51 Sectional pitching moment coefficient, normalized by tip dynamic pressure,
vs. azimuth at 15 m/s, 30◦ yaw. . . . . . . . . . . . . . . . . . . . . . . . . 100
52 Q-criterion iso-surface (Q = 0.1) at 15 m/s and 30◦ yaw, predicted using
HR-LES. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
53 Q-criterion iso-surface at 15 m/s and 0◦ yaw, computed using a compressible
algorithm [105]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
x
SUMMARY
environment. Even when the rotor is not stalled, the flow over the blades is dominated by
three-dimensional (3-D) effects. Stall is accompanied by massive flow separation and vortex
shedding over the suction surface of the blades. Under yawed conditions, dynamic stall
may be present as well. In all operating conditions, there is bluff-body shedding from the
turbine nacelle and support structure which interacts with the rotor wake. In addition, the
high aspect ratios of wind turbine blades make them very flexible, leading to substantial
aeroelastic deformation of the blades, altering the aerodynamics. Finally, when situated in
a wind farm, turbines must operate in the unsteady wake of upstream neighbors.
Though computational fluid dynamics (CFD) has made significant inroads as a research
tool, simple, inexpensive methods, such as blade element momentum (BEM) theory, are
still the workhorses in wind turbine design and aeroelasticity applications. These methods
tions with very limited empirical 3-D corrections. As a result, they are unable to accurately
predict rotor loads near the edges of the operating envelope. CFD methods make very few
limiting assumptions about the flowfield, and thus have much greater potential for predict-
ing these flows. In this work, a range of unstructured grid CFD techniques for predicting
wind turbine loads and aeroelasticity has been developed and applied to a wind turbine
configuration of interest.
First, a nearest neighbor search algorithm based on a k-dimensional tree data structure
was used to improve the computational efficiency of an approximate unsteady actuator blade
method. This method was then shown to predict root and tip vortex locations and strengths
similar to an overset method on the same background mesh, but without the computational
Eddy Simulation (HRLES) turbulence model, previously developed for structured grids,
xi
was extended to an unstructured framework. It was demonstrated to improve predictions
of unsteady loading and shedding frequency in massively separated cases. The sensitivity of
the model to highly stretched grid topologies was also explored. For aeroelastic predictions,
a methodology for tight coupling between an unstructured CFD solver and a computational
structural dynamics tool was developed. Due to the lack of experimental data pertaining
to a flexible turbine, the coupling algorithm was validated for a helicopter rotor, but the
method is sufficiently general that it can be immediately applied to a wind turbine when
suitable correlation data becomes available in the future. Finally, time-accurate overset
both RANS and HRLES turbulence models. The HRLES model was able to accurately pre-
dict rotor loads when stalled. In yawed flow, excellent correlations of mean blade loads with
experimental data were obtained across the span, and wake asymmetry and unsteadiness
xii
CHAPTER I
MOTIVATION
The United States is in need of alternative energy sources to replace current hydrocarbon
sources, including coal, natural gas, and petroleum. These resources are non-renewable and
emit greenhouse gases into the atmosphere. In the case of petroleum, a large portion of
the global supply comes from geopolitically unstable regions of the world. It is estimated
that the United States has sufficient wind resources to provide up to 20% of its energy [1].
Figure 1 shows a map of wind energy availability in the United States. There are currently
large stretches of land with high wind availability that are grossly underutilized. However,
there are a number of challenges, both technical and political, that must be overcome before
of power generated per dollar when compared to more traditional resources. The torque
output of a wind turbine, which is directly related to the power output, scales linearly
with the swept area of the rotor disc. Therefore, to increase the total generating capacity
of a wind farm, either larger turbine rotors or a larger number of turbines are needed.
Since such a large percentage of the total cost of a turbine is absorbed by the support
structure (i.e. the tower) and other parts that do not directly contribute to energy capture,
a smaller number of turbines with larger rotors is the most cost-effective strategy. On a
modern utility-scale turbine, rotor diameter can be on the order of 100 meters. As blade
length increases, aeroelastic effects become more important, contributing to fatigue that
shortens the life of a turbine. Furthermore, building blade and tower structures capable of
withstanding those loads increases weight and therefore cost of the turbine.
To better carry the structural loads of large turbines, blades with very thick airfoil sec-
tions have been proposed. The effect of these thick sections on aerodynamic performance
1
2
Figure 1: United States annual average wind power [31].
is difficult to predict using current tools, as they rely heavily on blade element momen-
tum theory (BEMT), which cannot accurately predict the response of blades that violate
thin airfoil assumptions. Furthermore, such methods model aerodynamics as basically two-
dimensional (2-D) phenomena, yet very large turbines with blunt airfoil sections exhibit a
great deal of three-dimensioanl (3-D) behavior in their flowfields. Analysis of such complex
fluid dynamics (CFD). Unlike BEMT methods, which neglect 3-D effects and aerodynamic
phenomena occurring off the blade surfaces, CFD models the complete aerodynamic envi-
the energy infrastructure. On a turbine, the major sources of broadband noise are inflow
turbulence, interaction of the turbulent boundary layer with the airfoil trailing edge, and
vortex shedding noise due to the finite thickness of the trailing edge [34]. However, there is
also very low frequency noise associated with vortex shedding from the tower [49], typically
on the order of 20 Hz. Low frequency noise is attenuated by the atmosphere less than the
higher frequency broadband noise from the blades. In a downwind turbine configuration,
this is only magnified by the interaction of the blades with the turbulent tower wake.
Recently, there have been reports of numerous health problems for people living close (less
than 1 km) to a wind farm, due to constant low frequency noise [24]. Aeroacoustic prediction
methods currently used in the wind industry tend to be empirical in nature, relying on
assumptions that may be incorrect for modern designs. Modern aeroacoustic methods
that use CFD flowfield data as their input stand to greatly improve acoustic predictions
since CFD proceeds directly from the underlying equations of fluid motion and relies on
fewer assumptions that limit its applicability. Since noise comes from infinitesimal pressure
fluctuations with magnitudes on the same order as error terms in CFD solutions, accurate
aeroacoustic predictions of this type require flowfield characterizations with accuracy that
3
CHAPTER II
BACKGROUND
Despite their relatively simple appearance, wind turbines operate in an aerodynamic envi-
ronment that is challenging to numerically model. Due to the atmospheric boundary layer,
there can be considerable variation in wind speed between the top and bottom of the rotor
disc. If the turbine is not facing directly into the wind, i.e. if there is yaw, the blade loads
will vary cyclically as they rotate. All wind turbines have some kind of yaw control, but
the wind direction can vary too quickly for the controller to maintain zero yaw. At high
simple, inexpensive methods based on blade element momentum (BEM) theory. These
methods provide basic insight into turbine flows, but only under the simplest conditions:
constant wind speed with zero yaw. BEM methods operate under what Leishman calls
the independence principle, in which the aerodynamics of each airfoil section along the
blades are computed independently of neighboring sections [45]. As a result BEM methods
completely neglect spanwise flow and other 3-D effects, which have been shown to result
in significant lift augmentation and stall delay, especially near the blade roots [83]. As a
result, traditional BEM methods, even when they incorporate some sort of 3-D correction,
typically under-predict torque. Designs based on such simulations can result in structures
Computational fluid dynamics techniques can mitigate many of the inaccurate simpli-
fying assumptions used in common wind turbine analysis methods. However, as will be
discussed shortly, CFD analyses have their own shortcomings. What follows is four broad
classes of CFD techniques that can offer improvements over current wind turbine design
tools. The first is actuator blade methods, which can provide physics-based characteriza-
tions of wind turbine wakes while reducing computational expense associated with modeling
4
the blades. The second is hybrid Reynolds Averaged Navier-Stokes / Large Eddy Simuation
(RANS/LES) methods, which can improve the prediction of unsteady, separated flows. The
third is overset methods, which introduce the ability to treat the relative motion between
rotor and its support structure. The final technique is coupling between CFD and a com-
putational structural dynamics methodology to model the aeroelastic response of the rotor
blades.
Actuator disc techniques are one approach to managing complexity and computational
expense in simulations of wind turbines. Actuator disc methods are based on momentum
theory and seek to model the effects of a rotor on the surrounding flowfield without modeling
the rotor itself. Since the flow over the rotor blades produces lift and drag forces on the
blades, there must be another force equal in magnitude and opposite in direction acting on
the flow. This reaction force can be included in the underlying CFD model in two ways. It
can be implemented like a special boundary condition, where a permeable boundary surface
is embedded in the mesh, with the flow velocity constrained to be continuous through the
surface while the pressure is discontinuous. Alternatively, those forces can be included
as body forces in the Navier-Stokes equations. The body force approach offers additional
flexibility over the pressure discontinuity. Since no surface is needed in the grid, the position
of the disc can change, allowing different choices of the tip path plane on the same grid. In
either case, the blades themselves are neglected. The pressure discontinuity or body forces
provide the same effects on the flowfield as rotating blades, but in a time-averaged sense.
Without modeling the blades directly, CFD simulations can either use fewer grid points,
speeding their execution compared to more detailed methods, or place more points in the
CFD simulations that employ actuator discs has seen widespread use in both the ro-
torcraft and wind industries. For a good review of actuator disc techniques, largely from
the perspective of structured grid CFD, see Reference [42]. O’Brien implemented unsteady
actuator disc and and blade techniques in the unstructured solver Fun3d and compared
5
those against fully overset techniques on several geometries, finding that an actuator disc
dramatically improved predictions of helicopter fuselage loads [68, 67]. Schweikhard [86]
implemented a steady-state actuator disc in another unstructured framework that was very
similar to that of O’Brien except that the boundary surface approach was used. While
this approach fixes the location of the rotor relative to other bodies in the flow, the use of
boundary surfaces allows for flat boundary layer-like cells to be used at the disc to capture
the strong gradients normal to the disc without excessively increasing node count in the
radial direction.
While actuator disc methods provide a good approximation of the influence of the rotor,
the fact that they lack discrete blades means that they model the rotor in an azimuthally
averaged fashion. This makes them unsuitable for cases with yaw error since the rotor wake
varies azimuthally, not just radially. Furthermore, though O’Brien showed that they can
capture the roll-up of two wing-like vortices in the near wake, they cannot capture discrete
tip vortices [67]. To work around this problem, Sørensen et al. [96] used an actuator-line
method, whereby the blades are modeled by lines along which body forces act. These body
forces are typically derived from a BEM method that extracts aerodynamic response from
tabulated airfoil data. In this manner, the unsteady effect on the flowfield by individual
blades can be included in analysis while still avoiding direct CFD modeling of the blade
surfaces and their associated boundary layers. This actuator line method was applied to a
500 kW Nordtank turbine and achieved good agreement with the experimental power curve
for pre-stall wind speeds. Above about 12 m/s, where that particular rotor is stalled, the
method over-predicted power because the airfoil data, lacking three-dimensional effects, is
The actuator line method has also been applied with good results by Mikkelsen [58], who
compared it against an axisymmetric actuator disc approach and used it to validate some
of the fundamental assumptions of traditional BEM methods. Whereas Sørensen et al. [96]
used a code that solves the Navier-Stokes equations in vorticity-velocity form, Mikkelsen
used the incompressible code EllipSys3D, which solves in pressure-velocity form. This
decision allows for the inclusion of solid boundaries like the turbine tower.
6
The EllipSys3D code was also used by Ivanell et al. [38], who evaluated the vortical
wake structures produced by the actuator line method. In that work, the actuator lines were
fixed in the grid, and the effects of blade rotation were applied via a boundary condition,
which allowed them to use an efficient steady-state formulation. While this significantly
decreases computational expense, it makes the method unsuitable for yawed cases. In
addition, the azimuthal boundary condition was also periodic, so a three-bladed rotor could
be modeled with a grid covering a 120◦ sector of the rotor disc. Again, this precludes the
method being applied to any case that does not have a periodic solution, exactly the cases
In addition to actuator discs, O’Brien also developed an actuator method that lay be-
tween the actuator disc and line concept. This method, termed actuator “blades,” was
evaluated on helicopter rotor-fuselage interaction problems [68, 67]. The key difference be-
tween actuator line and actuator blade methods is that in the former, each blade consists
of a single line of sources. To avoid discontinuities, each source’s loading is distributed over
multiple grid points using a “regularization function” that makes a source’s influence at a
distance r away from it scale with exp(r2 ). On the other hand, the actuator blade method of
O’Brien uses a rectangular array of sources for each blade, providing a continuous influence
without a regularization function. This should provide a more accurate local influence since
the local angle of attack (a key input to the underlying BEM model) can vary with chord.
Though they are undoubtedly easier to apply to complex geometries, unstructured actu-
ator methods like that of O’Brien [67] suffer from one drawback besides the usual computa-
tional overhead associated with unstructured grids. In order to apply body forces from the
actuator sources to the flow, it is necessary to know which grid node (or cell centroid for a
cell-centered code) is closest to each actuator source, which is continuously in motion with
respect to the inertial grid. This entails some sort of nearest neighbor search procedure. In
actuator disc cases or actuator line/blade cases in which the sources are fixed in the grid
(perhaps because the equations are being solved in a rotating frame), this search can be
performed as a one-time pre-processing step. In that case, any search algorithm will be
sufficient. However, if the sources move, the search must be repeated at each time step.
7
In the structured grid actuator methods discussed so far, the grid has a regular structure
in memory that mimics the structure in space, allowing one to efficiently search through
a range of grid indices. In an unstructured grid, a much more general search procedure is
manner, by looping over all the grid nodes to find the node associated with a single source,
and then repeating that loop for all the other sources.
Numerous other approximate techniques exist for predicting rotor flows, including vortex
lattice and vortex particle methods [20, 85]. These methods have the advantage of not
requiring a large volume mesh to discretize the flowfield and record the disturbances created
by bodies within them. Instead, the bodies shed particles that carry vorticity into the wake.
However, they often rely on idealized models of fluid flow that prevent them from being
Most fluid flows of engineering interest are turbulent, and wind turbines are no exception.
While numerous advances have been made in the numerical solution of the Navier-Stokes
equations, turbulent flows still present challenges for today’s methods. Turbulent flows are
characterized by a very wide range of scales in both time and space. Most of the kinetic
energy of a turbulent flow is stored in the large-scale structures of the flow. In contrast,
kinetic energy is dissipated as heat at the smallest scales. The size of the smallest length
scales, and therefore the maximum allowable CFD grid spacing to completely resolve all
turbulence, is inversely proportional to Re3/4 . So for a 3-D simulation, the number of grid
points must scale with Re9/4 [27]. Because such fine grids and small timesteps are not
practical with today’s computers, compromises must be made that approximate certain
Most CFD methods solve the Reynolds-averaged Navier-Stokes (RANS) equations. Es-
sentially, only the mean flow is solved on the computational mesh, and the turbulent physics
are replaced by closure models of varying sophistication. RANS methods include simple
8
Figure 2: Illustration of the resolution of turbulence scales by numerical techniques.
algebraic models like Baldwin-Lomax, one-equation models like Baldwin-Barth and Spalart-
Allmaras, and two-equation models such as k-ω and k-ω SST. These turbulence models tend
to be heuristic and rely on non-physical constants that are “tuned” to specific flows, such as
airfoils at low angles of attack. That is, the results from the model are iteratively compared
against experimental data, and the constants are adjusted until the computational results
agree with the experiment. While they perform well for those cases, they fail to accurately
predict unsteady flows dominated by viscous effects, as in the case of static and dynamic
Large-eddy simulation (LES) is a more physically accurate CFD technique than RANS,
in which the governing equations are mass-averaged and then filtered such that turbulence
eddies larger than the grid size are captured and turbulence with scales smaller than the
grid is still modeled. As Davidson notes [27], the rationale for this is that very little energy
actually cascades from the smaller scales back into the large scales, so approximations
made at the small scales should have few effects on the large scale flow features that are of
engineering interest. Strictly speaking, LES requires that the resolved scales extend into the
inertial sub-range. These scales, and those smaller, are more universal, meaning that their
physics depends less on the particular geometry. Such universal scales are more amenable
to modeling, since even when heuristics are used, they should be valid for a larger range of
flows. The drawback of LES is that in order to directly solve for the larger turbulent scales,
it requires much finer grids and smaller time steps than unsteady RANS (URANS). In most
engineering applications, LES remains far too computationally expensive for routine use.
Hybrid RANS/LES methods provide a way to achieve some of the advantages of LES
for separated and highly vortical flowfields while retaining the computational efficiency of
9
RANS. Baurle et al. [12] developed the idea that RANS and LES methods can be linearly
blended by some smooth function to form a hybrid model. This idea of blending was utilized
in 2006 by Sanchez-Rocha et al. [79], who used the k-ω SST RANS model as the basis for a
hybrid method within an existing LES code to resolve wall-bounded turbulence. Sanchez-
Rocha demonstrated this capability with simulations of a NACA 0015 airfoil in static and
an LES model into a RANS code to capture subgrid scales, so that where grid resolution
permits, LES-like results are obtained. On coarse grids that are usually only suitable for
RANS applications, hybrid methods of this variety can capture larger turbulent eddies in
the interior of the flow (away from boundaries), thus providing a better approximation of
These two approaches to hybrid RANS-LES, are different in the sense of the underlying
methodologies. LES codes are often explicit, with high order spatial and temporal schemes.
Care is taken to include most, if not all, of the fluctuating terms that appear when mass-
averaging the compressible Navier-Stokes equations. Within existing CFD (URANS) codes,
the spatial and temporal schemes are usually of lower order than LES codes. This is
especially true of unstructured codes, where spatial accuracy beyond second order is very
difficult to implement. The temporal integration is typically implicit to allow larger time
steps while maintaining the stability of the numerical scheme. Fluctuating terms appearing
be simplified so that they are less computationally expensive. Finally, the schemes may be
Detached Eddy Simulation (DES) is a common hybrid formulation, in which the turbu-
lence near walls is modeled with Menter’s k-ω SST [54] or Spalart-Allmaras (SA) [98] RANS
turbulence models. However, it must be noted that DES models are separate and distinct
from hybrid methods like those of Sanchez-Rocha in that they lack a dedicated subgrid-
scale model. Instead, the RANS equations perform “double-duty” as the LES model by
modifying a length scale used in the destruction terms [109, 106]. In the baseline SA model,
distance from the nearest wall is used as a characteristic length scale. In SA-DES, that
10
distance is replaced by a term proportional to the local grid spacing. This allows for a very
simple implementation and does not necessarily imply a negative impact on the validity of
the solution. In contrast, the method of Sanchez-Rocha has a separate LES model that is
linearly blended with a RANS model. A key disadvantage of DES methods is that great
care must be taken, either during grid generation or in the computation of the length scale,
to avoid grid-induced separation, in which the model transitions to LES too early, in bound-
ary layer regions of the mesh that are too coarse to support it. The Delayed DES (DDES)
model successfully addresses this shortcoming by supplying a limiter that depends on the
Recently, Vatsa and Lockhard investigated the application of hybrid turbulence models
to aeroacoustic applications [112]. The models included a DDES model that was slightly
modified to avoid spurious upstream eddy viscosity and the GT-HRLES model, which is
developed later in this work. In general, both hybrid models were found to accurately
predict mean surface pressures and surface pressure power spectral densities. Additionally,
results with the modified DDES model in structured and unstructured frameworks were
compared and found to be essentially identical. However, this did not indicate one way or
the other whether unstructured methods pose any special challenges for hybrid turbulence
models since in both cases a structured grid was used, though in the case of the unstructured
equivalence would use a mesh of tetrahedra and prisms, which is more representative of a
typical unstructured grid, that is carefully constructed to resemble the structured grid in
particular engineering application, the individual blades can be modeled, though at greater
computational expense.
In 1999, Duque et al. [29] employed the overset CFD solver Overflow to compute
the flow over the NREL Combined Experiment Phase II rotor. These early Navier-Stokes
11
results were all at zero yaw, but included the nacelle and tower. While the results were
the overset mesh. In 2003, Duque et al. [28] performed Navier-Stokes simulations on the
NREL Phase VI rotor and compared that CFD analysis with the rotorcraft comprehensive
code Camrad, which used various stall delay and dynamic stall models. The Navier-Stokes
computations were shown to more accurately predict stalled rotor performance and also
Due to the complexity and expense of a full Navier-Stokes solution to wind turbine
problems, various approximations have been used in CFD analyses. One approach is the
hybrid method of Xu and Sankar [115], in which the Navier-Stokes equations are solved
only in a small zone approximately 2 chords wide surrounding each blade. The rest of the
flowfield is treated with a significantly less expensive full potential method. A free wake
method is also used in the potential zone to model the wake formed by the tip vortices.
This method showed promise at low wind speeds when flow separation is minimal, but at
Instead of searching for reduced-order numerical methods, other researchers have cho-
sen to employ complete Navier-Stokes methods, but with drastically simplified models that
eliminate much of the geometry except a single blade. In 2004, Le Pape and Lecanu used
ONERA’s code elsA, a structured multi-block solver, to model the NREL Unsteady Aero-
dynamics Experiment in an upwind, zero-yaw configuration [43]. Rather than model the
entire rotor, they modeled a single blade, and as was common at the time, omitted the tower
and nacelle. Furthermore, they solved the Navier-Stokes equations in a rotating frame of
reference. Torque and thrust predictions in that study were in good agreement with the
experimental data at lower wind speeds, before the onset of stall. But at moderate speeds,
Though CFD has largely been perceived as too expensive for routine use on full turbines,
it has seen a great deal of use in analysis of individual turbine parts, such as advanced
airfoils. Stone et al. [103] used Overflow to evaluate the aerodynamic performance of
flatback airfoils, a new type of airfoil with a very blunt trailing edge designed to better
12
carry the structural loads of a large turbine. They compared the flatback performance, as
model, and a hybrid RANS-LES model, to that of the conventional airfoil. As expected,
the DES and HRLES models were able to capture the inherently 3-D physics of the airfoil
wake, unlike the S-A model. Given the great influence of 3-D and rotational effects on
the aerodynamics of even conventional rotors [83], this study highlights the importance of
Chao and van Dam [22] used Overflow to study the torque produced by a notional
flatback version of the NREL Phase VI rotor compared to the actual rotor. They found
that flatback airfoils are viable alternatives to conventional airfoils in large rotors in that
the torque is not seriously degraded compared to a conventional rotor at low to moderate
wind speeds. They also noted that, due to the low freestream velocities, low-Mach pre-
conditioning was essential to achieving accurate solutions when using a compressible flow
solver. However, due to limitations in the LMP method in Overflow at the time, they
rather than moving the rotor blades through a background mesh. While they obtained
useful results, this method cannot be applied to yawed rotors, since the flow over a yawed
Using an incompressible overset method, Zahle et al. [117] showed that CFD can capture
the unsteady physics associated with interactions between tower and rotor wakes. In that
study, a two-bladed NREL Phase VI rotor was set downwind of a simplified tower, without
a nacelle. The rotor lacked a physical connection to the tower and instead “hovered”
just upstream of it. However, the major physical phenomena of blade-wake interaction
were captured since the tower obstruction was present. Further accuracy was obtained by
including a simple ground plane. As is common in other work discussed earlier, agreement
with experimental performance data was reduced at higher wind speeds, an effect attributed
to turbulence modeling.
In a second study by Zahle et al. [118], the same incompressible overset method was
applied to an upwind rotor with inflow shear to emulate an atmospheric boundary layer.
13
They found that the inflow shear resulted in a lag in the blade loads compared to a case with
uniform inflow as the blades pass in front of the tower. Furthermore, the blades exhibited
dynamic stall effects in that region. Finally, they found that with the shear inflow, the
wake dissipates faster downstream of the rotor due to mixing between the upper and lower
One of the earliest unstructured CFD analyses of the wind turbine rotor was conducted
by Sezer-Udol and Long [87]. Their simulations of the NREL Phase VI rotor were strictly
inviscid, and blade motion was enabled by rotating the entire mesh, which prevented the in-
clusion of a tower. Though their inviscid method did capture some of the massive separation
present at 15 m/s, rotor thrust was in all cases nearly double the measured value.
Finally, Potsdam and Mavriplis [74] modeled the NREL Phase VI rotor using the un-
structured Navier-Stokes code Nsu3d. They modeled the rotor in a zero-yaw configuration
using both steady-state and time-accurate formulations, and a range of different surface
mesh topologies. They also investigated adaptive mesh refinement to better capture the
turbine wake. They found that performance results obtained with the unstructured code
compare favorably with those obtained with the structured overset solver Overflow, but
that due to the problem of high aspect ratio boundary layer cells, a problem shared by
all unstructured codes, unstructured grids may require more nodes for the same level of
accuracy. Using the Spalart-Allmaras turbulence model, the onset of stall was delayed and
the power over-predicted in the early stall regime, indicating the need for more advanced
turbulence models.
Coupling between computational fluid dynamics (CFD) and computational structural dy-
namics (CSD) tools has been used extensively in the rotorcraft community to predict un-
steady blade loads with greater accuracy than traditional comprehensive codes [77]. Sophis-
ticated structural models are often used in blade element-momentum (BEM) codes for wind
turbine design. Furthermore, Navier-Stokes CFD has become more common in recent years
for analysis of particular design points. However, the two are rarely used in combination.
14
Since there has been little use of CFD/CSD coupling methods in wind turbine applications,
this section largely focuses on the development of these methods for rotorcraft. Fortunately,
the coupling methods originally developed for rotorcraft analysis can be adapted for wind
turbines. However, there are some important differences in assumptions that need to be
addressed before applying rotorcraft coupling methodologies to wind turbines. These can
include non-uniform freestream velocity due to wind shear and turbulence; flexible tower ef-
fects, resulting in an oscillating hub position; and construction defects, which lead to rotors
with slight variations in structural and aerodynamic properties from one blade to another.
2.4.1 Definitions
Coupling refers to the simultaneous (or nearly simultaneous) solution of the equations de-
scribing the flow over the rotor and the deformation of the blades. If the two sets of
equations are not solved simultaneously, then the CFD and CSD tools that handle each set
must communicate in some manner. The CSD code computes time-accurate blade defor-
mations and provides those surface motions to the aerodynamic code, which uses them to
modify its boundary conditions. The CFD code computes the flow over the entire flowfield
and integrates those loads on the blade surfaces. It provides those loads to the CSD code,
which uses them as forcing functions for the structural model. There are two principle types
of coupling:
Loose coupling means that the fluid and structural equations are solved using separate
tools that communicate at some interval greater than the time step of one of the codes.
Common schemes have the two codes exchanging loads and motion once per rotor revolution
or once per blade passage (nb times per revolution, where nb is the number of blades). The
two codes might not, and often do not, use the same time step.
Tight coupling means that the two codes are still executed separately but they communi-
cate once per time step. Clearly, this can become computationally expensive, depending on
the code communication methods used (file I/O vs. library calls). Due to the computational
expense, loose coupling is usually favored in situations where the airloads and structural
deflections are expected to be periodic. Tight coupling is required in other situations, such
15
as in maneuvering flight.
A third possible type of coupling is intimate coupling or full coupling, in which the fluid
and structural equations are solved simultaneously, in the same solver [8]. Though such an
approach has the advantage of having no costs associated with communication of loads and
motion, this type of coupling is almost never used in practice for rotating systems. Intimate
coupling would require creating a dedicated analysis tool that would be extremely complex.
Due to the computational expense involved, the CFD half of the coupling problem has
always been the limiting factor. For this reason, the development of coupled CFD/CSD
methods has essentially tracked the development of CFD for routine aerospace analysis
tasks. First full potential methods were applied to isolated blades, then on full rotors.
Then the Euler equations were solved on isolated blades, then on rotors. More recently, full
In the early 1980’s, Johnson developed the comprehensive analysis tool Camrad. For
computational efficiency, it used lifting line aerodynamics and a prescribed wake model. In
1984, Tung et al. performed limited coupling between Camrad and the full potential code
Fpr in order to capture transonic effects on the advancing side of the rotor that are beyond
the scope of a lifting line model [110]. Because these transonic effects only exist near the
tip, and because the full potential computations were substantially more expensive than
other parts of the analysis, the full potential “zone” was limited to the tip region. In 1989,
boundary layer model to Fpr in order to more accurately determine viscous drag on the
advancing side.
A common feature of early coupling methods was that rather than deforming the blade
grids and thus mimicking the physical deformation, they relied instead on modifying the
boundary conditions used in the CFD solver. Typically, blade motions would be used to set
up a transpiration boundary condition in which flow was allowed normal to the unmoving
blade surfaces [18]. In addition, they avoided using CFD throughout the computational
16
domain, instead focusing on selected areas that were difficult to handle with lifting line
aerodynamics. Due to the limited coverage of the CFD zone, the rotor wake was excluded,
and its influence was accounted for using a varying inflow angle.
solver with a structural solver for hovering rotors [92]. In this work, a single blade was
structurally modeled as steady-state beam problem. The flowfield was computed using a
fourth order spatial algorithm with a compact stencil, which allowed for solutions using
40-50% fewer grid nodes than would have been required in a conventional second order
In the early 1990s, overset CFD methods gained popularity due to the ease with which
they could model complex real-world geometries. One of the first widely used overset
Navier-Stokes solvers was NASA’s Overflow. In its original form, it was capable only
of static overset simulations. In 1994, Overflow-d was released with the capability to
handle arbitrary grid motion. This innovation made it very popular for rotorcraft CFD.
Indeed, while rotorcraft CFD has by no means been restricted to any solver in particular, a
good overview of its development can be had by focusing on Overflow simulations alone.
Shortly after Overflow-d was developed, Bauchau and Ahmad [8] performed some
of the first examples of tight CFD/CSD coupling between Overflow-d and the flexible
multi-body dynamics code Dymore. Unlike other comprehensive analysis tools in existence
at that time (like Camrad) which operate in the frequency domain and rely on a modal
analysis of the structural model, Dymore works strictly in the time domain and integrates
the discretized equations of motion directly. Operating in the time domain is a requirement
for tight coupling since all CFD solvers also operate in the time domain. The work of
Bauchau and Ahmad is also distinguished from other aeroelastic work discussed here so
far in that they solved the thin-layer Navier-Stokes equations on a computational domain
that covered the entire rotor without reliance on a wake model. However, due to the large
computational expense, no rotor trim was performed; control angles from experimental data
were used instead. Even without trimming, their computed airloads showed clear differences
17
In 1997, another tight coupling analysis, that of Lee, Saberi, and Ormiston [44], high-
lighted some of the algorithmic issues that would be heavily investigated in coming years.
They tightly coupled the comprehensive analysis tool 2gchas with the full potential solver
Fpx. As in previous studies that applied a full potential solver rather than Euler or Navier-
Stokes CFD, they had only partial CFD coverage of the computational domain. Previous
loose coupling results had difficulty achieving convergence of the coupling process, and as
stated by Lee et al., “it was suggested that the loose coupling procedure was inadequate
when pitch moments/torsional dynamics were included.” Overcoming this difficulty was the
reason tight coupling was used for their analysis. Even so, they still experienced torsional
instabilities, proving that the coupling algorithm was not at fault. Instead, they found that
errors in the pressure distribution at the trailing edge of the blades contributed to negative
pitch damping.
In a typical coupling problem, the rotor collective and cyclic pitch controls are not known
a priori. In a procedure called trimming, the controls must be adjusted until the mean
integrated hub loads match certain target values. Typically, the collective and cyclic pitch
are adjusted to match thrust and rolling and pitching moments. Trimming creates particular
difficulties for the stability of a loose coupling procedure. Since coupling is performed
at some interval greater than the CFD time step, and the CFD loads are held constant
throughout a coupling iteration, the loads applied to the structural model are not able to
In the last few years, a particular loose coupling algorithm that addresses these problems
of poor trim and convergence has become the norm in CFD analysis of rotorcraft. In this
so-called delta trimming method, the loads applied to the structural model are actually the
sum of loads from a low-fidelity lifting line model and a delta load equal to the difference
between the lifting line and CFD loads [77, 26]. At coupling iteration k, the applied forces
18
and moments on the structural model, F k , are given by:
F k = FLL
k
+ ∆F k−1 (1)
k k−1 k−1
= FLL + FCF D − F LL (2)
k−1 k k−1
= FCF D + FLL − FLL (3)
FLL and FCF D are forces obtained from the lifting line and CFD models, respectively. As
the solution progresses, FLL should stop changing, so at convergence the applied load would
k−1
simply be F k = FCF D . Since the lifting line loads are free to change at any time step, they
One of the first studies to use the delta trimming method was that of Potsdam et al.
[77], in which Overflow-d was coupled with Camrad-ii and Rcas and applied to the
UH-60A rotor. CFD was used throughout the computational domain, and the blade grids
were deformed elastically, mimicking the physical deformation that occurs in flight. This is
in contrast to earlier studies that relied on transpiration boundary conditions on the blades
to include the effects of structural deformation. Motion and loads communication between
the codes was accomplished via the exchange of specially formatted files on disk (which has
since become the norm for loose coupling). These computational results were compared
against experimental airloads from the extensive UH-60 flight test database. Three forward
flight conditions were analyzed: the high speed case known as c8534, the low speed case
c8513, and the high thrust case c9017. For all cases, convergence of the trimming algorithm
The high speed case is less demanding from a CFD perspective because the wake, which
can be difficult to capture without very fine grids or high order algorithms, is convected
away from the rotor quickly enough that it has little effect on airloads. Good agreement
with experimental sectional normal force Cn M 2 was found, indicating the success of the full
Navier-Stokes approach. While the overall trend of the sectional pitching moments Cm M 2
was good, their magnitude was consistently over-predicted inboard and under-predicted
outboard. In addition, there was a 25o phase shift in the pitching moments.
The low speed case c8513 is dominated by blade-vortex interactions (BVI), which are
19
notoriously difficult to capture with CFD since most schemes rapidly diffuse the vortices.
Nevertheless, both the magnitude and phase of the normal force and pitching moments were
well-predicted, particularly the vibratory components with a natural frequency of 3/rev and
higher.
The high thrust case c9017 is at the edge of the UH-60 flight envelope and is dominated
by dynamic stall events. Phase shift effects were quite evident in this test case. Though
Potsdam et al. had no difficulty with convergence in this case, others have. Large changes
in control angles may not produce large changes in rotor loads since the blades may simply
Datta et al. [26] also applied the delta loads method to the UH-60 database by coupling
Turns and Umarc. For reasons of computational expense, the CFD method was somewhat
simpler than that of Potsdam et al. Turns only models a single blade; the effect of the
wake from other blades was included through a Bagai-Leishman free wake model. They
chose to focus on the prediction of vibratory airloads, which are particularly important for
both fatigue analysis of the rotor and airframe and passenger comfort (in civilian aircraft).
Datta et al. [26] experienced problems with both the magnitude and phase of airloads
the advancing side in the high speed case c8534, they explained that poor pitching moment
predictions are largely to blame. Pitching moment is the major driver in vibratory loads
due to its influence on the elastic torsion and hence the sectional angle of attack. Pitching
moment is much more sensitive to inaccuracies in the flow solution since it is computed
in large part by multiplying surface pressure forces by some moment arm, which can have
a magnifying effect. Despite these deficiencies, they found that the use of CFD yielded a
great deal of improvement over lifting line and free wake aerodynamics because CFD is able
to handle the transient shocks that occur on the advancing side in a way that other tools
simply cannot.
More recently, the problem of BVI has been tackled further by Lim et al. [46] using
loose coupling between Overflow-2 and Camrad-ii applied to the HART-I and HART-II
rotors and by Makinen et al. [50] using Overflow-2 and Rcas applied to HART-I. Again,
20
good convergence of the trim algorithm was observed. In both studies, results concerning
grid quality were presented. In the former, it was found that a “standard” grid of 20
million points showed great improvement in airloads prediction over a “coarse” grid with
only about 5 million points. However, a “fine” grid with over 100 million points showed
The studies discussed thus far all used algebraic or one equation turbulence models such
as Baldwin-Barth and Spalart-Allmaras. The work by Makinen et al. [50] went further in
using a newly developed HRLES turbulence model [88], which is a more physics-based
they were better able to take advantage of very fine off-body grids and capture more of the
All of the CFD/CSD coupling efforts discussed so far have used structured CFD meth-
ods. In principal, the CFD mesh type should have no bearing on the applicability of a
particular coupling methodology. In practice, unstructured grids are more difficult for cou-
pling because very few assumptions can be made about the mesh geometry and so more
rigorous procedures must be used for common tasks such as integrating airloads and de-
forming meshes. For example, when determining the loads to apply to the CSD model,
pressure and viscous forces are usually calculated on spanwise slices of the surface grid. In
a structured grid, slices are usually parallel to constant i, j, or k lines, making it straight-
forward to determine which surface faces contribute to the load on an airstation. With
an unstructured grid, a detailed search procedure is needed to identify such faces, either
through an exhaustive search of every boundary face or the use of some auxiliary data
Abras developed a number of important techniques for using the delta coupling method
of Potsdam [77] in an unstructured grid solver. Specifically, she coupled the unstructured
solver Fun3d with the flexible multibody dynamics code Dymore to predict UH-60 for-
ward flight cases [4, 3]. These simulations illustrate the extra care that must be taken in
an unstructured method. For example, in a structured solver the volume grid might be de-
formed by simply translating and rotating each spanwise grid plane, but this is not feasible
21
in an unstructured solver. Abras used a “spring analogy” to determine how the grid should
be deformed in response to the lifting line deflections provided by Dymore. She found
that to the deformed mesh from being excessively stretched, it was necessary to apply those
Biedron and Lee-Rausch investigated loose coupling with Fun3d and Camrad-ii, ap-
plied to the HART-II rotor [16]. In addition to demonstrating another variant of the loose
delta coupling technique, they performed a detailed study of the grid resolution and tem-
poral accuracy required to achieve acceptable correlations with experimental data. In par-
ticular, they found that resolution of the prominent blade-vortex interaction (BVI) events
that occur in the HART-II descent cases require much finer grid resolution than that used
In nearly all coupled rotorcraft simulations, it is assumed that while the blades may
bend and twist, their cross-sections remain fixed. In 2009, Thepvongs et al. [108] developed
a coupling algorithm intended for studying rotors with morphing airfoils. Instead of a
beam representation, the blades were modeled using a 3-D finite element model capable
of large deformations. The structural code included a finite-state aerodynamic model, and
coupled aeroelastic predictions compared favorably with predictions using only the lower
order aerodynamics.
22
CHAPTER III
This chapter contains descriptions of the basic methods used by the two CFD solvers used
in this work, Overflow and Fun3d, as well as the flexible multibody dynamics tool
Dymore.
3.1 Overflow
Overflow [19] is an overset structured grid CFD solver developed primarily by NASA.
The use of overset grids simplifies modeling of flows around complex geometries or moving
bodies since a given grid need only cover part of the geometry rather than the entire
implicit finite-difference formulation. By default, fluxes are computed using a 2nd order
accurate central difference scheme, though spatial accuracy as high as 4th order is available.
A Roe upwind scheme is also available for the inviscid fluxes. Unsteady solutions can be
1st or 2nd order accurate in time. To extend the range of Mach numbers over which
the compressible method can be used, unsteady low-Mach preconditioning (ULMP) [76]
is available. Overflow has been used successfully for fixed-wing, rotorcraft [5, 53, 47],
and wind turbine applications [29, 28]. A number of turbulence models are available in
Overflow, including k-ω SST and GT-HRLES, which was implemented as part of the
3.2 Fun3d
Fun3d is NASA Langley’s code for solving the Navier-Stokes equations on tetrahedral or
mixed element unstructured grids [17, 6]. It can solve both compressible and incompressible
flows, the latter computed using using Chorin’s method of artificial compressibility [23]. It
uses an implicit node-centered finite-volume formulation. Steady flows are solved using a
1st order backward Euler scheme, while time accurate schemes are solved using a 2nd order
23
backward differentiation formula (BDF). By default, the inviscid fluxes are evaluated using
Roe upwinding. The viscous fluxes are always evaluated using a scheme that is equivalent to
a central difference formulation. The resulting linear system is solved using a point-implicit
Overset grids can be used in Fun3d in conjunction with DiRTlib (Donor Receptor
Transaction Library) [62] and SUGGAR (Structured, Unstructured, and Generalized Grid
AssembleR) [63]. The primary use of overset methods in Fun3d is moving body problems
[15] since unstructured grids already provide a powerful tool for handling geometric com-
plexity. Similar to Overflow, Fun3d has been used for a very wide range of applications,
3.3 Dymore
Dymore is a multi-body finite element code developed by Bauchau and others [10]. Dy-
more contains a library of primitive elements, such as beams, springs, hydraulic dampers,
and various types of joints, for which the equations of motion can be stated exactly. A
model consists of one or more primitive elements connected so that the connections enforce
the boundary conditions for each element. All equations of motion are expressed in a single
inertial frame and are integrated in the time-domain, in contrast to other codes, which of-
ten express the equations in body-fixed frames and integrate in the frequency domain. The
formulation used in Dymore is more general, which makes it applicable across a broader
range of applications, rather than being restricted to a few cases (e.g. loose coupling). The
24
CHAPTER IV
O’Brien demonstrated that an actuator blade methodology can provide a reasonable ap-
proximation of the unsteady loads on a helicopter fuselage due to the rotor wake [67, 69].
As discussed earlier, these methods model the influence of the rotor on the flowfield by
adding source terms to the Navier-Stokes equations that act as body forces.
the new positions of the actuator sources, which rotate through the mesh in clouds that
reproduce the approximate blade planform as shown in Figure 4, are computed. Then
a search is done to find the nearest neighboring mesh vertex to each source. Since each
processor stores all of the actuator sources, but only a subset of the mesh, a given source
may lie outside the mesh partition. In that case, the source’s true nearest neighbor actually
lies on another partition. To deal with that situation, the processors exchange the distance
they computed from each source to its nearest vertex. If another processor has a smaller
distance, that source is marked as belonging to another partition, and its influence is not
included on that processor. This “multiple partition check” is handled with an efficient
MPI_Reduce operation. Following that check, the state variables at the vertex associated
with each source are used to computed the local effective angle of attack. That angle
of attack is then used in a blade element momentum (BEM) model to compute source
strengths. Finally those source strengths are added to the residual and Jacobian, and the
linear system is solved. A more complete discussion of this overall process can be found in
Ref. [67].
haustive search was used in which the distance between each source and each mesh vertex
actuator sources, and Nv is the number of mesh vertices resident on a particular processor.
25
Figure 3: Flowchart of the actuator blade solution algorithm. The box with a dashed
border, indicating the search portion, is the only part that was altered.
26
(a) Complete rotor (b) Close-up of tip
Figure 4: Actuator sources (blue) embedded in mesh. Each source must be associated with
the nearest mesh vertex.
In a typical case, Nv might range from 30,000 to 200,000, and Ns might be on the order
of 2,000 per blade. Thus, for some configurations, this search can become a very expensive
operation. Indeed, O’Brien found that the cost of an actuator blade solution using this
implementation can approach the cost of a solution with overset blades [69].
Nearest neighbor searches have been thoroughly studied in the computer science literature,
and many search algorithms exist that are faster than the exhaustive search described above,
the set of points being searched is recursively divided into smaller subsets. The divided
set may be represented by a tree-based data structure like an octree or a kd-tree. Though
kd-tree literally means “k-dimensional tree,” it usually refers to a specific type of binary
tree in which each level of the tree is split along one Cartesian direction, with the direction
alternating in an x,y,z fashion through the tree. Such trees are known to be among the best
data structures for nearest neighbor searches in two or three dimensions [13]. In general, a
kd-tree nearest neighbor query has O(log N ) complexity, where N is the number of nodes
27
4.1.1 Tree construction and nearest neighbor queries
at most mv mesh vertices, including the vertices’ array indices and their positions in space,
as well as pointers to left and right child nodes. Each tree node also has a split direction, d,
and a split point, ~s. All vertices in the left child have pd ≤ sd , and all vertices in the right
any depth in the tree. Since the mesh vertices are static, the tree is constructed just once,
at the beginning of the simulation. The process begins with a node being passed a list of
nv mesh vertices (at the beginning, all vertices are passed to the root node). If nv ≤ mv ,
all of the vertices are inserted into the node’s list, and control returns to the parent node.
If nv > mv , the vertices are sorted by their position in the split direction, d. Once sorted,
the lower half of the vertices are sent to the left child, and the upper half are sent to the
right child. The median point alone is inserted into the current node. The split direction
alternates through the tree, so that if a node is split in direction, d, its two children are
For example, consider the following list of six 2-D vertices: (4, 7), (9, 6), (5, 4), (2, 3),
(7, 2), (8, 1). Assume mv = 1, meaning that only one vertex may be stored in each tree
node, and that the initial split direction is x. First, the list is sorted in the x-direction.
The median of the list is (7, 2) and is inserted into the root node of the tree (Figure 5(a)).
That vertex becomes the split point for the root node. The nodes to the left of (7, 2) are
now sorted in the y-direction. The median of that sublist is (5, 4), which is inserted into
the left child of the root (Figure 5(b)). The only vertex to the left of (below) (5, 4) is (2, 3),
so it goes into the left child of the root’s left child (Figure 5(c)). Vertex (4, 7) is handled
similarly (Figure 5(d)). Then control moves to the right half of the original sublist: (8, 1)
and (9, 6). That sublist is sorted in the y-direction. The median is (9, 6), which is inserted
into the right child of the root, and then that child is split at that point (Figure 5(e)). The
only remaining node is inserted into the left (bottom) child of the root’s right child (Figure
5(f)).
28
(7, 2)
(7, 2)
(7, 2)
(5, 4)
(5, 4)
(7, 2)
(7, 2)
(5, 4)
(2, 3) (5, 4)
(7, 2)
(2, 3)
29
(4, 7)
(7, 2)
(5, 4)
(2, 3) (5, 4)
(7, 2)
(2, 3) (4, 7)
(4, 7)
(9, 6)
(7, 2)
(5, 4)
(2, 3) (5, 4) (9, 6)
(7, 2)
(2, 3) (4, 7)
(4, 7)
(9, 6)
(7, 2)
(5, 4)
(2, 3) (5, 4) (9, 6)
(7, 2) (8, 1)
(2, 3) (4, 7) (8, 1)
30
Searching a kd-tree for the vertex nearest to a query point, ~q, can begin at any node,
though typically it begins at the root node. The number of points stored in a node and its
simple exhaustive search is done over its vertices. That is, the distance squared between
q and p is computed for each vertex and then compared to the smallest distance found so
far. If it is smaller, it becomes the new “best” distance, and that vertex’s index is saved.
Control then returns to the parent node since terminal nodes have no children.
If the node has nv > mv , the distance squared between ~q and the one point actually
stored in the node is computed and compared to the minimum distance encountered so
far. If the query point lies to the left of the node’s split point (qd < sd ), the left child is
recursively searched. Otherwise, the right child is searched. After that search, the distance
from the query point to the plane separating the two children is computed. If it is smaller
than the minimum distance encountered so far, it is possible that the nearest vertex to the
query point lies just across that plane in the child that was not searched previously. In that
case the other child is searched recursively as well. Finally, control is returned to the parent
node.
For example, consider the example tree constructed earlier and a query point (1, 5). A
search for the query point’s nearest neighbor begins at the root node, with vertex (7, 2).
The distance squared from the query point to that vertex is 45 (Figure 6a). Since the query
point is to the left of the root node’s split point, we now search the root’s left child, which
contains the vertex (5, 4). The distance to that vertex is 17, which is the best encountered
so far (Figure 6b). Since the query point is to the “right” of (above) the split point, we
now search the right child, with vertex (4, 7). That distance is 13, which is better still than
the previous closest vertex (Figure 6c). Since this node has no children, we now check the
distance to the plane separating it from its sibling. The distance to the plane is 1, which
is less the the distance to the best vertex found so far (Figure 6d). Therefore, a vertex
lying just on the other side of the plane could be a nearer neighbor, and so we search the
sibling. The vertex contained in that sibling has a distance of 5 from the query point, and
so becomes the best candidate (Figure 6e). Now control returns to the parent node, with
31
vertex (5, 4). Both its children have been searched, so control returns to the root node. The
distance from the query point to the root node’s split plane is 36, which is greater than
the distance for the best candidate, and so the right child of the root node need not be
searched (Figure 6f). The nearest neighbor of the query point is (2, 3). Since the search
never descended into the right tree, the expensive distance function was never evaluated for
those vertices.
they are the total number of vertices contained in the tree, Nv ; the maximum number of
vertices contained in any one node, mv ; and the efficiency with which the distance function
is evaluated. The third factor is addressed by always using squared distances so as to avoid
computing square roots and by computing distances inline without a separate function call.
To determine how the search algorithm scales with the maximum number of vertices
allowed per tree node, a simple scaling study was done on a mesh with 154,082 vertices,
varying mv . Altering mv essentially changes the number of nodes in the kd-tree. Since
kd-tree searches scale logarithmically with the number of nodes, mv was changed by powers
of two. Figure 7 shows the ratio of the average cost of a kd-tree search to an exhaustive
search vs. mv . In this limited test, the best performance was obtained with mv set to either
undertaken on two similar grids. The grids contained about 2.4 and 4.5 million vertices,
respectively. Both grids contained no solid boundaries. Mesh points were clustered in a
region between the rotor disc and another disc about three rotor diameters downstream.
The number of vertices in each CPU’s tree was altered by varying the number of CPUs used
in each simulation. The number of vertices per processor ranged from 17,651 to 282,420.
For each of those node counts, simulations were conducted using the pre-existing exhaustive
search; a kd-tree search with mv = 128, previously identified as optimal; and a kd-tree search
with mv = 16, a reasonable but still sub-optimal value. The average search time, as well
32
(4, 7)
(1, 5)
(9, 6)
(7, 2)
dist = 45
(5, 4)
(2, 3) (5, 4) (9, 6)
(7, 2) (8, 1)
(2, 3) (4, 7) (8, 1)
(4, 7)
(1, 5)
(9, 6)
dist = 17 (7, 2)
(5, 4)
(2, 3) (5, 4) (9, 6)
(7, 2) (8, 1)
(2, 3) (4, 7) (8, 1)
dist = 13 (4, 7)
(1, 5) (9, 6)
(7, 2)
(5, 4)
(2, 3) (5, 4) (9, 6)
(7, 2) (8, 1)
(2, 3) (4, 7) (8, 1)
Figure 6: Graphical depiction of the algorithm for search a kd-tree wit mv = 1. Vertices
are represented by black dots, the query point by a black cross, and the current nearest
neighbor by an unfilled circle. Gray regions are not currently being searched.
33
(4, 7)
(1, 5)
(9, 6)
(7, 2)
dist = 1 < 13
(5, 4)
(2, 3) (5, 4) (9, 6)
(7, 2) (8, 1)
(2, 3) (4, 7) (8, 1)
(4, 7)
(1, 5)
(9, 6)
(7, 2)
dist = 5
(5, 4)
(5, 4) (9, 6)
(2, 3)
(7, 2) (8, 1)
(2, 3) (4, 7) (8, 1)
(4, 7)
(1, 5) dist = 36 > 5 (9, 6)
(7, 2)
(5, 4)
(5, 4) (9, 6)
(2, 3)
(7, 2) (8, 1)
(2, 3) (4, 7) (8, 1)
Figure 6: Graphical depiction of the algorithm for search a kd-tree wit mv = 1 (continued).
Vertices are represented by black dots, the query point by a black cross, and the current
nearest neighbor by an unfilled circle. Gray regions are not currently being searched.
34
Figure 7: Ratio of the cost of a kd-tree search to the cost of an exhaustive search vs. the
maximum number of vertices per CPU.
as the average total time (including activities not related to actuator blades) per time step
The kd-tree to exhaustive search cost ratios are shown in Figure 8. Clearly, when
the CPUs are very heavily loaded with a large number of vertices, the kd-tree is vastly
superior, with less than 15% the search cost of the exhaustive search (solid lines). There
are diminishing returns from the kd-tree as the number of vertices per CPU decreases (i.e.,
as the number of CPUs increases, since the total number of vertices is fixed). On the
larger mesh (grid 2), the kd-tree is 88% as expensive as the exhaustive search. Also, at
low numbers of vertices, the search procedure is overall a very small percentage of the total
time per time step. This occurs because the exhaustive search makes very efficient use of
The effect of mv , the maximum number of vertices per tree node, is striking. Allowing
more vertices per tree node improves performance of the kd-tree considerably, as can be seen
in the distance between circles and triangles in Figure 8. Increasing mv allows the kd-tree
algorithm to gain back some of the cache efficiency of an exhaustive search. Furthermore,
it reduces the maximum depth of the tree so that fewer recursive function calls are required
35
(a) Grid 1, with 2.4 million vertices (b) Grid 2, with 4.5 million vertices
Figure 8: Ratio of the cost of the kd-tree search to the cost of the exhaustive search. Solid
lines indicate the cost ratio for the search portion of a time step alone. Dashed lines indicate
the cost ratio for a complete time step.
to reach a terminal node. Indeed, as can be seen on the smaller of the two grids (Figure
8(a)), using 128 vertices per node rather than 16 ensures that the kd-tree search is always
In the Fortran implementation of the present search algorithm, a kd-tree node is defined
according to Listing 4.1. Since the two arrays holding the list of vertices are statically
allocated for speed, the memory usage for one kd-tree node is always:
In all applications considered here, k = 3. In a best-case scenario for memory usage, each
terminal node would hold exactly mv vertices, minimizing the total number of nodes. In a
worst-case scenario, each would hold only one vertex, giving the largest possible number of
nodes and essentially wasting 28(mv − 1) bytes per node. So memory usage will be in the
range:
Nv
[24 + 28mv ] < total memory < Nv [24 + 28mv ] bytes (5)
mv
Currently, memory usage is within acceptable bounds. Should it become necessary to reduce
memory usage, the algorithm could be modified to directly index the arrays holding the
mesh vertices rather than copying them (into the pos variable in Listing 4.1. This would
36
Listing 4.1: Fortran definition of a kd-tree node
type kdnode
integer :: dir ! splitting direction
integer :: npts ! no . points in or below this node
integer , dimension ( m_v ) :: index
! indices of the points
real ( dp ) , dimension (k , m_v ) :: pos
! positions of the points
type ( kdnode ) , pointer :: left , right ! children
end type
Actuator blades were used to simulate the NREL Phase VI wind turbine. A wind speed of
7 m/s was chosen since at higher speeds the flow over the blades begins to stall. Obviously,
the predictive capability of an actuator blade method will suffer when the flow is dominated
by viscous effects near solid boundaries or other phenomena, like stall, that depend greatly
on the details of the geometry involved. Each actuator blade had 2000 sources (100 in the
radial direction, and 20 along the chord, uniformly distributed) in keeping with O’Brien’s
recommendation [67] that the source array be finer than the grid in which it is embedded.
This source distribution prevents singularities in the actuator forces and keeps tip vortices
from being too diffuse. An incompressible formulation was applied as the freestream and tip
Mach numbers are very low, less than 0.1. Since no solid boundaries are present, the distance
function used in most turbulence models cannot be evaluated, and so the simulations were
laminar. The mesh included a total of 2.4 million vertices. An overset simulation with
two blades but without a tower was also conducted to compare the wake approximated by
actuator sources to the wake from actual rotating blades. The background mesh of the
overset simulation was identical to the mesh used in the actuator blade simulations. In
both cases, a time step equivalent to 1◦ of azimuth per step was used.
The actuator blade method originally implemented by O’Brien uses a linear approxi-
mation of the blade twist, which is well-suited to helicopter blades. However, wind turbine
blades often have very non-linear twist distributions; the NREL Phase VI blade is no ex-
ception. For that reason, the algorithm was modified to allow the user to supply the twist
37
distribution in a file. The twist distribution is interpolated to get an accurate blade twist
at each actuator source. Figure 9(a) illustrates the wake computed using a linear approxi-
mation of the blade twist. As the wake visualizations show, these approximations can lead
to very diffuse root and tip vortices. In addition, the sign of the apparent angle of attack,
which includes the blade twist, changes around 80% span, leading to tip-like effects further
inboard than expected. Finally, the linear distribution results in a non-physical vortex sheet
downstream of that 80% span location. Figure 9(b), which was computed using the actual
Figure 9(c) shows the overset solution. Away from the blades, the overset and actuator
solutions (with correct twist) are quite similar. The tip vortices predicted by the actuator
blades are indistinguishable from those predicted by overset blades. The overall wake deficit
is also well-captured. Naturally, the actuator blades can not mimic the starting vortex
captured in the overset solution, but that is a momentary transient effect. The “jet” of
higher speed flow through the center of the disk is somewhat more diffuse in the actuator
solution. This can be attributed to differing geometry near the hub. The inboard edge of
the actuator blades corresponds to the first radial station with an airfoil profile, 1.257 m
from the axis of rotation. The actual blades have a cylindrical root attachment starting
0.508 m from the hub followed by a gradual transition to the first airfoil section at 1.257
m. This simplification was necessary because the underlying BEM model of the actuator
method assumes a linear airfoil-like relationship between angle of attack and lift, which will
not hold for a cylindrical section. In contrast, the overset blades match the actual NREL
model blades. In addition to being closer to the hub, the cylindrical root sheds a more
The trajectories of the vortices after leaving the blade root and tip provide another
useful comparison of the actuator and overset blade methods. Figure 10 shows the axial
and radial components of those trajectories. From the axial component, it is clear that
the actuator and overset blade methods convect tip vortices downstream at nearly the same
speed. On the other hand, the actuator blade method convects the root vortices faster. This
is qualitatively evident from the contour plots of w/V∞ in Figure 9, where the actuator blade
38
(a) Actuator blades with linear approximation of actual twist. Note the non-physical vortex sheet.
Figure 9: Q = 0.05 iso-surfaces and axial velocity, w/V∞ , along a centerline plane at V∞ = 7
m/s, zero yaw. Q is normalized by R and Utip . All cases ran four revolutions at 1◦ of azimuth
per step.
39
results have a larger high speed core. The principal difference in the radial component of
the root vortex trajectory is an inboard offset of about 0.6 m for the actuator blade solution.
This is mostly due to the omission of the cylinder-to-airfoil transition in the actuator model,
The actuator blade method was originally developed for predicting unsteady loads on
helicopter fuselages. An analogous problem for wind turbines is predicting loads on a turbine
tower and nacelle, but those loads are of little engineering interest compared with the inertial
loads of the spinning blades. However, since the actuator blade method captures the gross
features of the turbine wake, it has potential for use in predicting interactions between
multiple turbines in a wind farm. An upstream rotor with actuator blades (or its steady-
state counterpart, the actuator disc) can be applied to predict the power loss experienced
by a downstream turbine in its wake. Actuator blades could also approximate an unsteady
upstream wake feeding into a high fidelity overset simulation downstream. In that manner,
blade vortex interaction between a turbine and its upstream neighbors could be predicted.
The principal difficulty in such a simulation would be the computational expense of running
the simulation long enough to allow the wake to convect from one turbine to the next and
the requirement of fine enough grid resolution between them to avoid dissipating the wake.
This latter requirement can be alleviated by the application of adaptive grids [37, 114, 91]
40
(a) Axial position, z of vortices vs. vortex age.
Figure 10: Trajectories of root and tip vortices for both overset and actuator blades (with
actual twist). The radial separation between the overset and actuator root vortices is
approximately equal to the distance between the overset and actuator blade roots.
41
CHAPTER V
HYBRID RANS/LES
Even in zero-yaw states, wind turbine flowfields have very prominent unsteady, 3-D effects,
including massive separations. These flow features cannot be accurately resolved using
model and its extension to an unstructured methodology is discussed. The model is then
The instantaneous Navier-Stokes equations for compressible flows can be expressed in tensor
form as:
∂ρ ∂
+ (ρuj ) = 0 (6)
∂t ∂xj
∂ ∂ ∂p ∂σji
(ρui ) + (ρui uj ) + − =0
∂t ∂xj ∂xi ∂xj
∂ 1 ∂ 1 ∂qj ∂
ρ e + ui ui + ρuj h + ui ui + − (ui σij ) = 0
∂t 2 ∂xj 2 ∂xj ∂xj
For flows relevant to common aerodynamic applications, several assumptions about the
flow can be made to affect closure of these equations. The flows of interest are assumed
comprised of a Newtonian fluid consisting of a monotonic gas with isotropic viscosity. Thus
the perfect gas law and the Boussinesq constitutive relation can be applied.
The difference between RANS and LES first appears in the averaging technique applied
to the Navier-Stokes equations. RANS equations apply Reynolds averaging where the prim-
~ , p, e) are separated into mean (f ) and fluctuating (f 0 ) components, then
itive variables (ρ, V
averaged over a finite period of time much greater than the turbulent fluctuation frequency,
so that the mean value of a single fluctuating variable will be zero. The mean value of some
42
multiplied fluctuating variables will remain non-zero, and these correlated values must be
to the incompressible Navier-Stokes equations to obtain closure for the viscous turbulence.
For LES simulations, density, heat transfer and pressure remain as Reynolds-averaged val-
ues, while the remainder of the pertinent variables are decomposed using a Favre-averaging
that accounts for compressibility effects, and yields mass-averaged (fe) and fluctuating (f 00 )
components. It should be noted that the fluctuating terms that arise in the Favre and
Reynolds averaging processes are not identical and therefore are typically assigned different
notations.
formulated as:
∂ρ ∂
+ (ρuej ) = 0
∂t ∂xj
∂ ∂ ∂p ∂σij ∂ 00 00
(ρuei ) + (ρuei uej ) = − + − ρui uj
∂t ∂xj ∂xi ∂xj ∂xj
∂ 1 ∂ 1 ∂ ∂ 1 00 00
ρ ee + uei uei + ρuej h + uei uei
e = (−qLj + uei σij ) − ρu u
∂t 2 ∂xj 2 ∂xj ∂t 2 i i
∂ 1
+ −uej ρu00i u00j + uej ρu00i u00i − ρu00j h00
∂xj 2
1
+ σji u00i − ρu00j u00i u00i (7)
2
there arises a Reynolds-stress tensor that requires closure, either in the specific Reynolds-
averaged form (τij = −u0i u0i ) or the compressible Favre-averaged form (ρτij = −ρu00i u00i ).
The turbulent kinetic energy can then be defined in its specific formulation k = 12 u0i u0i or
conserved form ρk = 12 ρu00i u00i , respectively. Favre-averaging also gives rise to the turbulent
∂u00
heat flux (qT i = ρu00i h00 ) and the rate of turbulent dissipation (ρ = σji ∂xij ). For flows
up through the low supersonic regime, the molecular diffusion and turbulent transport
terms (σji u00i − ρu00j 21 u00i u00i ) are typically ignored [113], and this practice is continued for this
analysis. The additional equations through which these terms are resolved or closed gives
43
∂ρ ∂
+ (ρuej ) = 0
∂t ∂xj
∂ ∂ ∂p ∂σij ∂
(ρuei ) + (ρuei uej ) = − + − (ρτij )
∂t ∂xj ∂xi ∂xj ∂xj
∂ 1 ∂ 1 ∂ ∂
ρ ee + uei uei + ρuej h + uei uei
e = (−qLj + uei σij ) − (ρk)
∂t 2 ∂xj 2 ∂xj ∂t
∂
+ −uej ρτij + uej ρk − qTj (8)
∂xj
If symmetry is assumed, the Reynolds-stress tensor yields six unknowns that are ap-
proximated using models about the behavior of the fluctuating correlations, u00i u00j . These
approximations yield the set of RANS turbulence models, ranging from algebraic to two-
equation techniques. It was previously noted that the current practice is to assume the
Boussinesq approximation, which can be utilized to relate the fluctuations to an eddy vis-
cosity, µT :
1 ∂f
uk 2
ρτij = 2µT Sij − δij − ρkδij (9)
3 ∂xk 3
Similarly, the turbulent heat flux vector can be related to the eddy viscosity, µT , via pro-
µT ∂ e h µT cp ∂ Te
qT i = − =− (10)
P rT ∂ xej P rT ∂ xej
that introduces the turbulent Prandtl number, P rT , which can be either constant or vari-
able, depending on the application. Finally, the rate of turbulent dissipation can be ex-
pressed as
00 2 00
ρ = µ 2Sji Sij − ukk uii (11)
3
In addition to the Favre-averaging, the concept of LES is based on the direct capture
of the large turbulence eddies as part of the solution of the Favre-averaged Navier-Stokes
equations, relegating the smaller turbulent eddies to be modeled. This process is based on
the view that the larger turbulence eddies contribute significantly to the Reynolds-stress
tensor, while the smaller eddies are less significant. In order to separate these effects, in
addition to the averaging process, the variables in the equation of motion should also be
filtered (typically referred to as Favre-filtering) to obtain the small or subgrid scale (sgs)
44
turbulence. An excellent discussion of these filtering techniques can be found in Wilcox [113].
Near the surface of the configuration undergoing simulation, the turbulent eddy scales
reduce significantly, requiring in LES a very refined grid that increases the computational
resources beyond the reach of most engineering applications. As attached boundary layer
The correlation of these turbulence terms requiring closure to viscosity permits the RANS
simulation. The information exchange occurs via the turbulent kinetic energy, k. In this
work, the RANS turbulence model chosen to effect this closure is the Menter k-ω SST
turbulence model [55], based on its success in prior CFD applications, for example [93, 88,
78].
The Menter k-ω SST turbulence model resolves two differential equations that describe
the turbulent kinetic energy, k, as well as an approximation for the length scale based on
the dissipation per unit turbulent kinetic energy, ω. These equations are given by:
∂ ∂ rans ∂ui ∗ ∂ ∂k
(ρk) + (ρuj k) = τij − β ρωk + (µ + σk µT ) (12)
∂t ∂xj ∂xj ∂xj ∂xj
∂ ∂ γρ rans ∂ui 2 ∂ ∂ω
(ρω) + (ρuj ω) = τ − βρω + (µ + σω µT )
∂t ∂xj µT ij ∂xj ∂xj ∂xj
1 ∂k ∂ω
+ 2(1 − F2 )ρσω2 (13)
ω ∂xj ∂xj
where the rans superscript is used to denote the use of the Reynolds-averaged Reynolds-
∂ui
stress tensor. Menter [55] indicates that the production terms (τijrans ∂xj
) can be modeled
wx )2 + (vx − uy )2 .
The LES turbulent kinetic energy equation to obtain the subgrid scale data is one
45
Baurle et al. [12] have demonstrated that RANS and LES methods can be linearly
merged to form a hybrid model. Speziale [101] proposed an extension to this via the
Reynolds-stress tensor. Thus the RANS equations of motion and kinetic energy equation
∂ ~ ∂ ~ = ∂ (G ~ src + ∂ (G
~ trans ) + G ~ hybrid ) + G
~ hybrid
(E) + (uej E) (15)
∂t ∂xj ∂xj ∂xj Ttrans Tsrc
~ = {ρ, ρuej , ρE, ρk}, and the right hand side of the equation consists of the original
where E
~ trans ) and source (G
transport (G ~ src ) vectors excluding the fluctuating turbulence terms,
~ Ttrans and G
which have been formulated into new vectors, G ~ Tsrc that will be hybridized.
~ hybrid = F G
G ~ sgs .
~ rans + (1 − F )G (16)
T T T
Further details of the development of this hybridization technique can be found in publica-
The first term on the right-hand side of Eq. 14 is the k sgs production term. If this
term is not included in the hybridization, so that all production of turbulent kinetic energy
comes from the corresponding term in Eq. 12, the resulting model is very similar to a DES
model. In a DES model, the underlying RANS model functions as an SGS model in regions
of the grid with sufficient resolution. The RANS destruction terms are simply modified to
depend on some characteristic length scale [109]. The fact that the model discussed here
is hybridized with a separate SGS model means that it is not a DES model. However, if
the choice is made not to hybridize production, the hybridized destruction term behaves
similarly to the destruction term within a DES model based on the k-ω SST RANS model
[113]. Henceforth, the term hybrid RANS-LES (HR-LES) will refer to the full hybridization,
while hybrid RANS-DES (HR-DES) will be used to refer to a form where production is not
hybridized.
46
5.3 Unstructured Implementation
When implementing the HRLES scheme into any existing code, it is necessary to nondi-
mensionalize the turbulence model equations in accordance with the code conventions. For
∂ ∂ ∂xj 1 ∂ ∂ ∂ ∂t â∞ ∂
= = , = = (18)
∂ x̂j ∂xj ∂ x̂j L̂ ∂x ∂ t̂ ∂t ∂ t̂ L̂ ∂t
where the caret symbol represents a dimensional variable. The effect of this nondimen-
sionalization on the governing equations is illustrated below with the k and ω turbulence
equations. The Boussinesq approximation of Eq. 9 was used to form the production term
in Eq. 20.
∂ rans ∂ rans M ∞ µT 2 Re ∗ rans ∂ ∂k
(ρk )+ (ρuj k )= Ω − β ρωk + (µ + σk µT ) (19)
∂t ∂xj Re ρ M∞ ∂xj ∂xj
∂ sgs ∂ sgs 2µT M∞ 1 ∂uk 2 sgs ∂ui
(ρk ) + (ρuj k ) = Sij − δij − k δij
∂t ∂xj ρ Re 3 ∂xk 3 ∂xj
sgs3/2 sgs
sgs
k ∂ µ µ ∂k
+ C ρ + + (20)
∆ ∂xj Pr P rt ∂xj
∂ ∂ M∞ Re Re 1 ∂k ∂ω
(ρω) + (ρuj ω) = γρΩ2 − βρω 2 + 2(1 − F2 ) ρσω2
∂t ∂xj Re M∞ M∞ ω ∂xj ∂xj
∂ ∂ω
+ (µ + σω µT ) (21)
∂xj ∂xj
at solid walls is simply k = k sgs = 0. Since the dissipation equation is not blended, the
boundary condition on ω is that of the k-ω SST model. With the terms non-dimensionalized
47
where y is the distance from the nearest solid wall. Since this distance is zero along a
viscous boundary, it is instead set to the normal distance from the wall of the nearest node
Evaluations of a circular cylinder were utilized to verify the new turbulence method and
examine its impact on highly separated flows. The first series of evaluations compared the
ability of the unstructured hybrid RANS-LES method to capture the flow physics with
respect to RANS, structured hybrid RANS-LES, and LES results. Additional evaluations
applied the model to grid systems of differing element topologies to determine the behavior
The initial evaluation of the hybrid RANS-LES method simulates the flow of air over a
circular cylinder at a Mach number of 0.2 and a diameter-based Reynolds number of 3900
for standard sea-level conditions. The k-ω SST and HR-DES models (non-hybridized pro-
duction terms) were used to simulate this case. The grid about each cylindrical section
is a structured O-mesh that was converted to form an unstructured hexahedral grid. The
grid included 200 nodes in the wrap-around direction with 139 points expanding radially
outward with a 10% stretching ratio, yielding a y + < 0.02. While the wake was somewhat
coarse due to the uniform spacing of the grid as seen in Fig. 11, it was instrumental in
verifying the implementation of the model, allowing a direct comparison with the structured
approximately 200 steps per shedding cycle, assuming a Strouhal number of 0.21.
Figure 12 shows instantaneous contours of vorticity magnitude about the 2-D circular
cylinder using RANS (k-ω SST) and HRLES models. Even in two dimensions, HRLES cap-
tures the physics associated with the vortex shedding much more accurately when compared
to the baseline k-ω SST RANS results that have been run on an identical grid and with the
same numerical options. The RANS solution shows that the vortex shedding process has
been completely smoothed out by the model, while the vortex wake of the hybrid RANS-LES
48
(a) Complete grid (b) Close to surface
solution includes features that are more consistent with the physical assumptions. This is
borne out when the values of the turbulent kinetic energy are observed, as in Fig. 13, where
the turbulent kinetic energy shows a smeared wake region with no periodicity. Figure 14 is
a graphical representation demonstrating the blending of the two models. Here, values near
zero indicate LES-dominated regions and values near one indicate RANS-dominant regions.
Near the surface and attached regions, the blending function is close to one, indicating that
the RANS model is dominant. In the wake and aft of the separation point, the function
approaches zero, denoting that the flow field turbulence characteristics are dominated by
The computed Strouhal number for the 2-D hybrid RANS-LES simulation is 0.25, com-
pared with the experimental [70] value of 0.215 ± 0.005. The RANS simulation did not shed
vorticity, so it was not possible to compute a Strouhal number. The separation point pre-
dicted by the hybrid method is 86.6◦ , while the RANS method predicts separation at 85.1◦ ,
both close to the experimentally determined location of 86.0◦ ± 2◦ , which was extracted
from data obtained at ReD = 5000 [95]. The drag coefficient for hybrid RANS-LES simu-
lation is computed to be 1.5, which is significantly higher than the experimental[40] value
of 0.99 ± 0.05, but is in line with the value of 1.65 obtained by similar LES simulations[40]
computed also in two dimensions. The k-ω SST simulation resulted in a drag coefficient of
49
(a) k-ω SST (b) Hybrid RANS-LES
Figure 13: Turbulent kinetic energy about the 2-D circular cylinder.
50
Figure 14: Instantaneous contours of the blending function of Eq. 17 for the 2-D circular
cylinder. Lighter areas are RANS-dominated, and darker areas are LES-dominated.
0.887.
A second set of circular cylinder evaluations was evaluated on three different grid systems.
These evaluations use fully unstructured grids that were generated as a two-dimensional
mesh of triangle elements and quadrilateral elements (the latter in the boundary layer) and
then extruded in the spanwise direction to form prismatic elements. This strategy permits
a straightforward variation of the spanwise resolution, which has been demonstrated [94] to
be an important factor in accurate CFD solutions of separated flows. The first grid had 51
planes spaced evenly over a 4-diameter span, the second had 101 planes over 4 diameters,
and the third had 101 planes over 8 diameters, yielding the same number of nodes as the
second grid, but with larger cell aspect ratios. A radial cross-section of the grid, shown
in Fig. 15, contains 123,652 nodes, 203,492 triangular elements, and 21,363 quadrilateral
elements. Normal spacing was such that y + << 1. A refined nondimensional time step of
∆ta∞ /D = 0.025 provided approximately 1000 time steps per vortex shedding cycle. In
While hybrid RANS-LES was earlier shown to yield some improvement over k-ω SST
in characterization of the flowfield about a 2-D cylinder, the improvement is even more
51
(a) (b)
criterion, sometimes called Q criterion, for the 3-D cylinder at Re = 3900 on the three grids
The definition of Q is such that it is only positive near vortices. In flow visualizations,
this property can be used to filter out the “sheets” of high vorticity that emanate from
turbulent boundary layers, revealing vortices. While the k-ω SST model captures periodic
vortex shedding in three dimensions where it did not in two, those vortices are essentially
2-D across the span (Fig. 16(a)). In contrast, the hybrid RANS-LES results on all grids
show substantial spanwise variation and a much more irregular wake (Fig. 16(b-d)). Since
k-ω SST yields a result that could have been obtained on a substantially reduced grid
(whatever the method requires for a 2-D simulation), the majority of the grid nodes have
Table 1 delineates the various statistics for the cylinder predictions, including mean
drag coefficient, Strouhal number, and separation location. Strouhal number is calculated
from the frequency spectrum of the fluctuating lift. Separation location is given in degrees
over the circumference of the cylinder from the leading edge stagnation point to the point
52
(a) k-ω SST, 4 diameters, 101 planes (b) HR-LES, 4 diameters, 51 planes
(c) HR-LES, 4 diameters, 101 planes (d) HR-LES, 8 diameters, 101 planes
Figure 16: Iso-surfaces of Q criterion about a 3-D circular cylinder with varying grid reso-
lution.
53
Table 1: Predicted circular cylinder characteristics for various turbulence methods and grids
with number of spanwise planes, Nz , and spanwise extent, Z. Separation location is given in
degrees of azimuth from the leading edge stagnation point. Experimental Strouhal number
is from Reference [70], and separation location is from Reference [95] at ReD = 5000. LES
data is from Reference [40].
Nz Z Turbulence Mean Strouhal Separation
model CD no. location
– – Exp. 0.99 ± 0.05 0.215 ± 0.005 86 ± 2◦
101 4D k-ω SST 1.456 0.213 98.4◦
2 – HR-LES 1.5 0.25 86.6◦
51 4D HR-LES 0.971 0.216 85.8◦
101 4D HR-LES 0.919 0.217 84.3◦
101 8D HR-LES 0.939 0.217 84.7◦
48 πD LES 1.04 0.210 88.0◦
where skin friction along the cylinder centerline drops to zero. All 3-D cases predict the
the 3-D k-ω SST simulation yields a drag coefficient over 1.4, close to the result of a 2-D
HR-LES simulation. With 3-D HR-LES, drag is predicted much closer to the experimental
value, though it is still slightly under-predicted on all grids. All HR-LES simulations predict
the separation location within the experimental bounds, while k-ω SST predicts it over 10◦
farther aft.
Figure 17 shows mean pressure coefficient along the cylinder centerline, from the lead-
ing edge stagnation point (θ = 0) to 180◦ opposite. There, it is clear that the HR-LES
simulations better predict drag because they do not over-predict base suction to the extent
that the k-ω SST simulation does. Figure 18 shows the correlation between computed and
measured centerline Cp , with any points on a diagonal line with a 45◦ slope being in perfect
agreement with the measured data. The zoomed in region in Figure 18(b) indicates that
the grid with a diameter of 4D and 51 spanwise planes has the best correlation once the
flow is separated.
Unlike their structured grid counterparts, which only allow hexahedral cells, unstruc-
tured CFD solvers may allow a variety of cell types. In structured formulations, very high
cell aspect ratios are quite common in directions where flow gradients are expected to be
small. However, in unstructured frameworks, tetrahedral cells of very high aspect ratio,
54
Figure 17: Mean pressure coefficient along the centerline for the 3-D circular cylinder on
several grids with both RANS and HR-LES turbulence models. LES data is from Reference
[40], and experimental data is from Reference [65].
Figure 18: Correlation of computed mean pressure coefficient with measured value.
55
(a) Tetrahedral boundary layer cells (b) Prismatic boundary layer cells
Figure 19: Iso-vorticity contours around a 3-D cylinder, computed using the HR-LES model
on grids with different cell types in the boundary layer.
which are commonly seen in boundary layers, pose special problems. Figures 19(a) and (b)
depict iso-surfaces of vorticity magnitude on the 3-D cylinder using tetrahedral and pris-
matic boundary layer cells. The tetrahedral boundary layer solution indicates the presence
of numerical errors in the form of spurious vorticity upstream of the leading edge. These
errors are not present when the boundary layer is modeled with prismatic cells.
Node-centered solvers typically use an edge-based scheme in which fluxes are evaluated
along the edges of the primal grid. Since each primal edge corresponds to a face in the dual
mesh, this is approximately equivalent to evaluating fluxes across the dual faces. In highly
stretched tetrahedral cells, some dual faces and primal edges will be nearly parallel, yielding
a very poor approximation of gradients across those faces. In turn, this reduces the accuracy
of the reconstruction scheme, and errors propagate through the solution. Therefore, when
applying this hybrid RANS-LES model, care should be taken to use prismatic cells in the
boundary layer.
Profiles of the streamwise velocity, u/U∞ , at seven downstream wake traverse locations
on the mesh with 101 planes over four diameters are shown in Figure 20. These profiles
have been averaged in time and in space, in the spanwise direction. Experimental data
56
from Reference [48] as well as LES data of Kravchenko and Moin [40] is also included. In
the far wake, HR-LES slightly over-predicts the velocity deficit, but captures the spreading
of the wake observed in the experimental data. At the position closest to the wall, at
x/D = 0.58, HR-LES predicts the correct trough-like profile. At the first position in the
recirculation region, x/D = 1.06, HR-LES predicts a velocity profile that looks similar to
those closer to the cylinder, continuing to produce a relatively thin shear layer. In general,
in the recirculation region, the HR-LES profile “lags” in space behind the experimental and
LES data, with the profile spreading farther downstream. From Table 1, it is clear that
these discrepancies in the wake do not lead to large errors in drag or shedding frequency.
The errors in Figure 20 can be attributed in some part to grid topology. As described
earlier, the mesh was originally generated in two dimensions and then extruded in the span-
wise direction, which results in cells that have aspect ratios as high as 6 in the wake outside
the boundary layer and as high as 40 in the farfield. The hybrid RANS-LES destruction
term uses (cell volume)1/3 as a measure of the local length scale, which implicitly assumes
a nearly isotropic grid. This topology consideration is distinct from the earlier discussion of
boundary layer cell type in that the issue is not one of gradient reconstruction, but rather
one of reasonably estimating the length scale representing the largest eddy that can be
To test the effect of grid topology, a naturally-developing 3-D grid was created. Instead
of a defined number of spanwise points, the cells were allowed to expand in the spanwise
direction at the same rate as the other two directions. So while the surface resolution was
similar to the original mesh, the total node count was substantially lower, only 5 million.
Figure 21 shows a spanwise slice of the two grids near the four upstream traverses. Wake
velocity profiles at the seven traverses are shown again in Figure 22. The x/D = 1.54
and 2.02 locations show substantial improvement over the RANS predictions, with the
former attaining more of the expected V-shaped profile, though still spreading slower than
Another potential cause of the discrepancies in the velocity profiles in Figures 20 and 22
is the sub-critical Reynolds number of these simulations. ReD = 3900 was chosen to allow
57
Figure 20: u/U∞ at seven locations in the wake of a cylinder at Re = 3900 using a spanwise
extruded grid.
58
(a) Extruded mesh, 101 spanwise planes over 4 diam-
eters
Figure 21: Spanwise slices of the original mesh and the 3-D mesh used to evaluate the effect
of cell stretching. Red lines indicate wake traverse locations.
59
Figure 22: u/U∞ at seven locations in the wake of a cylinder at Re = 3900 using a fully
3-D grid.
60
correlation of HR-LES results with the LES work of Kravchenko and Moin [40], but at that
actually occurs in the shear layers separating the recirculation region from the outer flow
[81]. Yet, it is assumed that the flow is fully turbulent in these simulations. A dedicated
LES solver would better predict transition, provided the grid was fine enough to resolve the
instability waves. Those waves would then propagate upstream an influence the flow near
the wall. With a hybrid RANS-LES model, some improvement can be expected since due
to the elliptical nature of the governing equations at this Mach number, information from
the wake can propagate upstream. However, since the RANS model is still used at the wall,
instabilities captured in the wake by the LES model will be averaged by the RANS model at
the wall. As a result, HRLES methods still require a dedicated transition model to achieve
the best results at sub-critical Reynolds numbers. Furthermore, Kravchenko and Moin [40]
note that this case is very sensitive to disturbances such as the freestream turbulence level.
They do not specify the freestream turbulence intensity in their LES simuations, but in the
√
present HRLES simulations, it was set at k/U∞ = 4.7 × 10−4 , which may be lower than
Finally, the coefficients of the LES portion of the method examined in this work are
constant. Sanchez-Rocha [80] found that using constant coefficients could lead to artificial
turbulence dissipation. If the wake turbulence is dissipated, there is less mixing across
shear layers, possibly leading to the abrupt, square velocity profile seen at x/D = 1.54 in
Figure 20. These results may be improved using a fully dynamic implementation where the
61
CHAPTER VI
As new wind turbine rotors get larger, novel techniques, including lightweight composite
materials, are required to reduce stress on the blades and drivetrain. Flatback airfoils,
which have very blunt trailing edges have been proposed for this purpose because their
thick cross-section can bear those loads with less structural material [71]. However, due to
their blunt cross-section, increased strength comes at the price of increased drag [102].
The hybrid RANS-LES model was used to model the flow over a DU97 flatback airfoil,
which includes a 10%c thick trailing edge. The simulagion conditions were set to mirror, as
closely as possible, wind tunnel data obtained for this airfoil by Berg and Zayas [14]. All
cases were simulated using a compressible formulation with a Reynolds number of 3 million
based on chord and a Mach number of 0.2. The latter was chosen to be low enough that
compressibility effects are negligible and high enough that preconditioning is not necessary.
A nondimensional time step of ∆t = 0.005 was used to give approximately 500 steps per
Two variants of the hybrid RANS-LES turbulence model, denoted as HR-DES and HR-
LES here, were examined. In HR-DES, as discussed earlier, only the destruction term of
is hybridized, resulting in a model very similar to Spalart’s DES model. In HR-LES, both
The results for this case vary significantly with grid topology. Three different grids were
used:
Detailed analysis of the results on these grids is presented in the following sections.
62
(a) Immediate vicinity of airfoil
Figure 23: Grids used for DU97 simulations with wind tunnel walls. These 2-D grids were
extruded in the spanwise direction.
Because the wind tunnel configuration has substantial blockage—an airfoil with thickness
of about 0.4c in a tunnel with a height of 2c—simulations were first attempted with a grid
that included the top and bottom wind tunnel walls. This grid, shown in Figure 23, was
generated using Solidmesh [33] and AFLR2 [52, 51] and has a built-in angle of attack of 10◦ .
The grid was generated in two dimensions, and then, to create a 3-D grid, the mesh was
duplicated 33 times across a span equal to 0.5c or 5h, where h is the trailing edge thickness.
A no-slip boundary condition was used on the airfoil surface, symmetry planes on the side
walls, and farfield boundary conditions on the inflow and outflow planes.
Figure 24 shows iso-surfaces of the Q criterion for both k-ω SST and hybrid RANS-LES
simulations. It is clear that the k-ω SST results show very little vortex shedding from the
blunt trailing edge. The HR-DES results show several vortices in the wake. However, these
vortices are essentially 2-D structures. In comparison with the Overflow results of Stone
[104], which show clearly 3-D structures, this seems to indicate that the grid here is too
63
(a) k-ω SST (b) HR-DES
Figure 24: Isosurfaces of Qc/â∞ = 1.012 around the DU97 airfoil in the wind tunnel at
α = 10◦
.
coarse in the spanwise direction. A similar conclusion can be drawn from Figure 25 which
shows vorticity magnitude at midspan. Nevertheless, the hybrid RANS-LES results show
an obviously unsteady flow, with a line of vortices convecting downstream. In contrast, the
k-ω SST results, show a pair of stationary vortices just behind the trailing edge, yielding a
Table 2 provides various statistics for the flatback simulations, including mean and RMS
lift and drag, as well as Strouhal number. Both k-ω SST and HR-DES yield lift significantly
higher than the experimental value. Drag is also higher in the case of HR-DES, which is
consistent with the low spanwise resolution. Interpretation of Strouhal number for the two
cases is less straightforward. For both, the mode with the highest amplitude occurs at
St = 0.088. However, the spectra plotted in Figure 26 show that the HR-DES case has
a large secondary mode at St = 0.15, which is much closer to the experimental value.
There are several reasons why the integrated lift is so poorly predicted on grid 1. The
primary reason is grid quality. Tangential and normal spacing is very fine at the trailing
edge, with 330 nodes across the blunt edge. The comparatively large spanwise spacing
(≈ 0.0156c) means that cell aspect ratios are quite large at the trailing edge—on the order
64
(a) k-ω SST (b) HR-DES
Figure 25: Vorticity magnitude at midspan for the DU97 airfoil in wind tunnel at α = 10◦
.
Table 2: Time-averages and standard deviation of lift and drag for the DU97 flatback airfoil
at α = 10◦ . Experimental results were obtained at a nominal α = 11◦ . Overflow results
are duplicated from [104]. FUN denotes Fun3d, and OF denotes Overflow.
Grid Turb. model Code C̄L C̄D CL0 CD0 St
– Experiment – 1.57 ± 0.13 0.055 ± 0.005 – – 0.24 ± 0.01
65
(a) k-ω SST (b) HR-DES
Figure 26: Normalized frequency spectrum of the lift history for the DU97 case. The dual
peaks in (b) indicate a secondary shedding mode.
of 100—which was earlier demonstrated to present problems with this hybrid RANS-LES
formulation. To make matters worse, the grid cells, which were originally all prisms or
hexahedra (due to extrusion), were split into tetrahedra in a pre-processing step. This was
done to reduce CPU time since, at the time of these simulations, the pure-tetrahedra path
in Fun3d was nearly twice as fast as the more general mixed-element path. Unfortunately,
splitting these already fine aspect ratio cells into tetrahedra generated elements with very
small angles. Maximum face angles on the dual mesh were over 179◦ .
In the experimental configuration that grid 1 attempts to emulate, the wind tunnel walls
were actually flexible and porous “acoustic windows” made of Kevlar, which allowed sound
to reach microphones. That porosity is not accounted for in the boundary conditions used
here, nor is the flexing of the walls, which has never been quantified. So the CFD boundaries
which were intended to be more realistic than free boundaries may have actually introduced
were conducted using an overset structured grid without wind tunnel walls. The grid,
shown in Figure 27, was originally generated by Stone [104] for use with Overflow. There
is a fine O-grid surrounding the airfoil. A series of telescoping Cartesian meshes then extend
66
(a) Off-body Cartesian grids (b) Near-body O-grid
Figure 27: Structured overset on DU97 airfoil. Figure and grids taken from Ref. [104].
to the farfield. Each 2-D spanwise plane has 218,650 nodes. To use this grid in Fun3d,
the grid was converted to an unstructured hexahedral mesh using Suggar++ [64]. Unlike
grid 1, the hexahedral cells in this grid were not converted to tetrahedra.
In Table 2, grid 2 corresponds to a 2-D version of this structured overset mesh. Those
results, obtained by Stone [104] with Overflow are presented for comparison purposes.
Grid 3 is a 3-D version with 33 planes on a 0.5c span, for a total of about 7.2 million
nodes. On this grid, mean lift is within the experimental range for both codes, on both
2-D and 3-D grids. The drag coefficient with Fun3d on the 3-D grid is also on the edge
of the experimental bounds. Finally, on grid 3, the Strouhal number for lift oscillation is
somewhat higher than on grid 1, though still not close to that predicted by Overflow on
Iso-surfaces of the Q-criterion on grid 3, shown in Figure 28, indicate that HRLES(2)
captures a significantly more unsteady wake. The RANS solution is essentially 2-D, while
the large rolling vortices shedding from the HRLES(2) solution twist, varying their cross-
section as the move dowstream. In addition, near the trailing edge, streamwise braids
appear between rollers, a flow feature not visible in HRLES(1) results on grid 1 in Figure
24(b).
67
(a) k-ω SST (b) HRLES(2)
Figure 28: Q criterion iso-surfaces around the overset DU97 airfoil with free boundaries
(grid 3).
68
CHAPTER VII
The high aspect ratio of modern wind turbine blades makes them very flexible. Due to the
sufficient confidence in their predictions, particularly in loading scenarios near the edges
of the operation envelope. As a result, wind turbine structures are usually over-designed,
leading to increases in weight and cost. Furthermore, the difficulty in predicting aerody-
namic loads on stalled blades is a major factor in the industry trend away from mechanically
Though the underlying structural model in typical aeroelastic codes may be quite so-
phisticated, the aerodynamic models are usually based on some variant of blade element
momentum theory [60, 11]. However, advancements over the last two decades in aeroelas-
tic predictions for helicopter applications have demonstrated that using CFD to provide
unsteady loads to the structural model dramatically improves loads predictions [77, 26, 3].
Since the aerodynamic methodology has fewer simplifications of the underlying physics, it
there are no such datasets for a wind turbine with non-negligible aeroelastic effects. The
NREL Unsteady Aerodynamics Experiment, a very detailed dataset discussed earlier, in-
volved a turbine with relatively short, rigid blades. Therefore, the validation of a CFD/CSD
coupling process for a helicopter rotor is presented in this chapter. This process is suffi-
ciently general that few modifications will be required to apply it to a wind turbine.
The test case chosen to validate the coupling methodology is the HART-II rotor test [116],
which was conducted in 2002 to study the effect of higher harmonic pitch control on rotor
69
Table 3: Geometric parameters for the HART-II rotor.
Rotor radius, R 2m
Rotor chord, c 0.121 m
Solidity, σ 0.077
Pre-cone angle 2.5◦
Actual shaft tilt, αshaf t 5.4◦ aft
Effective shaft tilt, αef f 4.5◦ aft
Rotor speed, Ω 108.9 rad/s
noise and vibration. The rotor was situated on a model fuselage, which was at the end of
a long sting. Various geometric parameters for the rotor are given in Table 3. To emulate
descending flight, the shaft was tilted aft by 5.4 degrees by tilting the entire fuselage/rotor
assembly. This has been calculated to correspond to an effective shaft tilt of 4.5 degrees aft
when wind tunnel walls are not present. As discussed previously, a number of researchers
have demonstrated loose CFD/CSD coupling with the HART-II rotor [46, 50, 16].
The HART-II rotor lacks flap and lead-lag hinges between the hub and the blades. The
blades attach directly to the hub, and the inboard portion of the blade, which has an
elliptical cross-section and is much stiffer than the outboard portions, is allowed to bend
elastically to absorb some of the bending moment that would otherwise be transferred
to the hub. The Dymore model used in these simulations, shown in Figure 29, has a
number of simplifications compared to the actual rotor. First, most of the hub hardware is
omitted. The blades simply attach to a revolute joint at the hub axis. While the HART
blade is physically a single piece, here each blade is modeled as two beams: a relatively
stiff inboard “flex beam” and a more flexible outboard main blade beam. The flex beam is
constructed from a single third-order finite element. The main blade consists of eight third-
order elements. Figure 30 shows the relationship between the number of main blade elements
and the first torsion frequency. With eight finite elements, the first torsion frequency can
number of elements used in the model since Dymore’s performance scales with O(Ne m2 ),
70
Figure 29: Hub of the HART Dymore model.
where Ne is the number of elements, and m is the bandwidth of the stiffness matrix [9].
Since adding more one-dimensional beam elements does not change m, performance scales
linearly with Ne .
The pitch link is also omitted since there is a great deal of uncertainty as to its structural
properties. To approximate the effect of the control system stiffness on the blade response,
a torsional spring connects the flex beam and the main blade beam. In order for the
torsional spring to provide the best possible approximation of the control system stiffness,
the spring constant must be adjusted until the natural frequency associated with the first
torsional mode matches measured values. This “tuning” process resulted in a stiffness of
1632.4 Nm/rad, yielding the target natural frequency of 419.57 rad/s. Figures 31 and 32
show a fan plot and the first six mode shapes at the nominal frequency, respectively. The
computational model follows the trend of the measured frequencies, and the mode shapes
As stated in Table 3, the HART blades had a small pre-cone angle of 2.5◦ , which
was included in the Dymore model in the reference configuration. This means that the
deflections output by Dymore are relative to the pre-coned state. The aft shaft tilt was
applied to the model via the farfield velocity used by the aerodynamic interface, which
mirrors the manner in which shaft tilt was applied to the CFD model by setting the angle
71
Figure 30: Convergence of the 1st torsion frequency with the number of finite elements in
the blade beam.
Table 4: Dimensions of each component of the composite mesh. All cells are tetrahedra.
Component Nodes Boundary faces Cells Grid generator
Blades (each of 4) 1,156,735 114, 302 6,731,961 VGRID
Fuselage/background 8,971,420 63, 026 52,733,053 VGRID
Total 13,598,360 520, 234 79,660,897
of attack to 4.5◦ .
The mesh used for these simulations (Figure 33) included a total of 13.6 million nodes. The
relevant mesh statistics are summarized in Table 4. This mesh is identical to that used
by Biedron and Lee-Rausch in 2008, and as described in Reference [16], this mesh was the
The CFD portion of the coupled simulation followed as closely as possible the work of
Biedron and Lee-Rausch [16], in order to provide a correlation of the loosely coupled analysis
prior to tight coupling. A timestep corresponding to one degree of azimuthal rotation was
used, and the number of Newton subiterations was dynamically determined by a temporal
error controller [111], insuring that at each time step, residuals were always reduced to
less than 5% of an estimate of the temporal error. On average, using the error controller
72
Figure 31: Fan plot for the HART Dymore model.
resulted in about 30 Newton subiterations per step. To correlate with Reference [16], the
The goal of this work is to demonstrate a tight coupling method. However, this presents
some difficulties when trimming the rotor. During loose coupling, the CSD solver is provided
loads at all the airstations for a single revolution of the rotor. It then enters a “reference”
phase in which those loads are applied to the CSD model until the motion has become
periodic, usually after about 50 revolutions. Then the CSD trimmer enters a “perturbation”
phase in which each of the three pitch controls is perturbed by a small amount one at a
time. The perturbation of each control lasts for 5-10 revolutions. After perturbing all three
controls, the change in hub loads resulting from each perturbation is used to construct
an inverse Jacobian matrix. Then, that inverse Jacobian is used to compute a new set
of controls, and the CSD solver enters the “simulation” phase. Finally, it is allowed to
73
Figure 32: First six mode shapes for the HART Dymore model.
74
(a) Surface grid
Figure 33: 13.6 million node composite mesh used for HART-II simulations.
75
run until the loads again become periodic, about 100 revolutions. This procedure is often
termed an auto-pilot trimmer. Though over 150 revolutions may have been simulated in
total, the CSD solver is typically quite efficient and requires no further inputs from the CFD
solver, so this is not an unacceptable computational burden. For example, one auto-pilot
trim cycle in Dymore usually requires about one hour, depending on various factors, such
In tight coupling, the same trim procedure described above would be extremely expensive
since the CFD solver would have perform as many revolutions as the CSD solver. A number
of alternative trim algorithms have been proposed to deal with this situation, but those are
still an area of ongoing research [37]. For this effort, trim is computed with a conventional
loose delta coupling method until convergence has been reached, after which the simulation
is switched to a tight coupling algorithm. The initial loose coupling procedure (for trimming)
follows that used by Abras [3] and Biedron and Lee-Rausch [16]. The only difference is that
a quasi-steady trimmer was used in place of the auto-pilot. With a quasi-steady trimmer,
the user supplies an inverse Jacobian matrix (perhaps computed in a single auto-pilot run),
and that matrix is used to continually adjust the controls to drive the mean hub loads to
After switching to tight coupling, the controls are kept fixed. Provided the initial loose
coupling converged, a correctly functioning tight coupling method should result in minimal
change in the solution during its operation even though the elastic deformations can change.
7.3 Results
7.3.1 Validation of the coupled model
The loose coupling phase was allowed to converge via five coupling iterations. The “zeroth”
iteration was run for two revolutions, and subsequent iterations each ran for half a revolu-
tion. It should be noted that for a rotor with nb blades, a coupling iteration need only last
1/nb revolutions. At that point, the airloads from all the blades can be combined to create
a map of airloads over an entire revolution. However, as the blade controls undergo a step
change at the beginning of each coupling iteration leading to transients in airloads, running
76
Figure 34: Convergence of blade pitch controls with loose coupling iterations.
2/nb revolutions allows the airloads to become quasi-periodic state before applying them to
Figure 34 shows the convergence of blade collective and cyclic pitch, θ0 , θ1c , and θ1s ,
during the initial loosely coupled trim phase. The controls change very little after the
second iteration, though as will be shown later, the airloads continue to change. Only the
Dymore portion of the sixth iteration was simulated. The controls from the sixth iteration
were then fixed for the later tight coupling phase. The final controls are given in Table 5,
along with controls computed via Fun3d/Camrad coupling from Reference [16].
The convergence of airloads at 87% span is shown in Figure 35. Prior work with the
HART rotor shows that many predictions, including airloads and structural loads, have finite
offsets from the measured values. Since the fluctuating component is of greater engineering
interest than the mean value, Figure 35(c) shows pitching moment with the mean compo-
nent removed to show the fluctuating component only. It is clear that by the fifth loose
coupling iteration, the airloads have nearly converged on their final values. The primary
77
Table 6: Mean values subtracted from structural moments in second column of Figure 36.
Flap, Nm Torsion, Nm Lag, Nm
r/R = 0.17 r/R = 0.33 r/R = 0.17
BL measured -9.27 -2.90 -10.26
Fun3d/Dymore -3.05 -5.17 -64.49
Fun3d/Camrad -3.36 -5.18 -63.18
differences between the ultimate and penultimate iterations are observed for the normal
force predictions in the first and fourth quadrants, which feature prominent blade-vortex
interactions. Since the controls remain relatively constant after the second iteration, it can
be concluded that most of the adjustments made by Dymore after that point are elastic
In an effort to further establish the validity of the structural model, structural moments
from the present simulation were compared against measured moments and those computed
by Biedron and Lee-Rausch [16]. The present loosely coupled simulations capture the
character and magnitude of the flap moment better than Fun3d/Camrad. The torsion
and lag moments in Figure 36 are nearly identical, and Dymore has the same difficulty as
Fun3d/Camrad in capturing the lag moment, which is very lightly damped. For reasons
discussed earlier, the second column of Figure 36 shows the structural loads with their
A comparison of Fun3d/Dymore airloads from the final loose coupling iteration with
those predicted by Fun3d/Camrad is provided in Figure 37. The two coupling method-
ologies provide similar results at all azimuths, with Fun3d/Dymore capturing the BVI
When a tightly coupled solution is initialized from a previously converged loosely coupled
solution, the computed elastic deformations and airloads should not change.
Tip torsion, flap, and lag predicted after switching to tight coupling are shown in Figure
38, along with the motions from the final iteration of loose coupling. There is very little
change in torsion and flap, although there is a small decrease in flap near 180◦ azimuth.
78
(a) Normal force, Cn M 2
Figure 35: Convergence of airloads at 87% span for the loosely coupled trim phase. For
measured pitching moment, the mean value is -0.00258; for iteration 5, the mean pitching
moment is -0.00513.
79
(a) Flap moment at r/R = 0.17 (b) Flap moment, mean removed
(c) Torsion moment at r/R = 0.33 (d) Torsion moment, mean removed
(e) Lag moment at r/R = 0.17 (f) Lag moment, mean removed
80
(a) Normal force, Cn M 2 (b) Pitching moment, Cm M 2 , with mean removed
Figure 37: Comparison of airloads at 87% span computed using Fun3d/Dymore and
Fun3d/Camrad in Reference [16]. Mean pitching moments are -0.00258, -0.00424, and
-0.00513 for measured, Fun3d/Camrad, and Fun3d/Dymore, respectively.
Airloads at 87% span, shown in Figure 39, also show minimal change, particularly in pitching
moment. Due to the small change in flapping motion at 180◦ , there is a small change in
normal force. There are other small variations in normal force in the first quadrant BVI
events. Since the tightly coupled solution moves closer to the measured data, it can be
surmised that the initial loosely coupled solution was not completely converged. Otherwise,
The mean measured thrust is T = 3300 N. The mean measured rolling and pitching
moments are Mx = 20 N-m (positive for rolling to the left) and My = −20 N-m (positive
for nose-up pitching). At the end of loose iteration five, the computed mean hub loads were
T = 3326 N, Mx = 11.9 N-m, and My = −29.2 N-m. After a revolution of tight coupling,
the hub loads are nearly the same, with T = 3328, Mx = 10.2 N-m, and My = −10.7
N-m. Though the change in pitching moment as a fraction of the target value is large,
instantaneous pitching moment varies from -200 N-m to 160 N-m, so the change in the
81
(a) Elastic torsion (b) Flap (c) Lead-lag
Figure 38: Tip deflections predicted in tight coupling and the final iteration of loose cou-
pling. Bars on the measured data indicate blade-to-blade variation.
Iso-surfaces of the second invariant of the velocity gradient tensor, also known as the
“Q criterion”, are shown in Figure 40. Q criterion, defined earlier in Equation 23, has the
property that it is only positive near vortices. As such, it is a useful tool for distinguishing
between sheets of high vorticity and actual vortices. The roll-up of the individual tip vortices
into a single horseshoe vortex is clearly visible in these visualization, though the horseshoe
dissipates just after of the rotor disc after exiting the refined region of the background grid.
Just before 90◦ and just after 270◦ are regions where the blade tip can be observed to pass
through several tip vortices in rapid succession, leading to the blade-vortex interactions
discussed previously. In Figure 40(a), it is also clear that the system of tip vortices resides
primarily within the plane of the rotor, as one would expect since the rotor is tilted aft to
7.4 Performance
The flow solution was run on a Cray XT5 supercomputer. Each cluster node had two
2.3 GHz quad-core CPUs, for a total of eight cores per node. Four hundred cores were
used for the flow solution. A single core was used for hole-cutting with the other seven
cores on that node sitting idle to increase the memory available to the hole-cutting process.
Each revolution required 7.3 hours of wall time. A total of 5.5 revolutions were simulated,
82
(a) Normal force, Cn M 2
Figure 39: Sectional normal force and pitching moment 87% span, comparing loose and
tight coupling.
83
(a) (b)
including the final revolution of tight coupling, resulting in a total wall time of about 84
hours or 34,272 CPU hours. During tight coupling, Dymore had a negligible impact on
CPU time and memory usage (less than 1% each). Dymore required about 45 minutes
to trim each loose coupling iteration on a single processor. The Suggar hole-cutting and
overset assembly process, which runs as a single process on the first MPI rank, takes 25% of
the total wall clock time in both loose and tight coupling. A single overset assembly process
is currently used because the mesh partitioning scheme in Fun3d, which yields compact
partitions that minimize communication (and therefore maximize scalability in the flow
84
CHAPTER VIII
Currently, the most sophisticated technique to accurately model rotating systems, such as
a wind turbine, is an overset CFD simulation. Such a simulation can capture the unsteady
physics associated with the movement of the blades as well as the effect the tower and
nacelle have on the vortical wake. This chapter presents results from simulations of an
upwind, two-bladed turbine with the goal of predicting rotor performance as accurately as
possible.
The NREL Phase VI Unsteady Aerodynamics Experiment, completed in 2000 [35], was
wind tunnel at the NASA Ames Full Scale Aerodynamic Complex. The 5-m blades were
instrumented with a large number of pressure taps and strain gauges to record unsteady
blade loads and structural responses. Numerous operating conditions were tested, includ-
ing both upwind and downwind rotor configurations at wind speeds ranging from 5 to 25
m/s and yaw angles from 0 to 30 degrees. The purpose of the tests was to build a dataset
for use in studying the complex aerodynamic interactions of a wind turbine and for im-
proving computational prediction of structural loads. Even though modern turbines are
always pitch-regulated, the Phase VI rotor has become a de facto standard for evaluating
Following completion of the wind tunnel tests, NREL initiated a “blind” comparison
[90], in which researchers were invited to use their codes to predict the turbine performance
at each operating conditions. Many different types of prediction methodologies were used by
the participants, including blade element methods, free wake methods, and three different
CFD methods. Though the Phase VI turbine is relatively small by the standards of a
85
Figure 41: NREL Phase VI turbine mounted in 80 × 120’ tunnel.
larger machines. These include 3-D, unsteady flow over the blades, stall at higher wind
speeds, and stall-delay due to rotational augmentation [84, 82, 83]. In the blind study,
there was very little agreement in the different performance and loads predictions, even in
In the present work, selected cases in the upwind baseline configuration (Sequence S),
were simulated using an unstructured overset methodology. The nominal blade tip pitch in
this case was 3 degrees. The leading edge pressure probes normally present on the blades
were removed and the flow was allowed to freely transition to turbulence (as opposed to
being tripped). This sequence of cases was chosen in part to allow correlation with the
S with sequence M, in which the boundary layers were tripped near the leading edge,
shows that at the wind speeds simulated here, transition effects are minimal, which is
important since the computational methodology assumes fully turbulent flow. As will be
discussed later, however, the issue of transition is not quite as straightforward as whether
In an overset simulation, the geometry is broken into smaller components, each of which
is meshed independently of one another, and then reassembled and permitted to overlap
86
Table 7: Dimensions of each component of the composite NREL Phase VI mesh. Boundary
layer cells are triangular prisms; non-boundary layer cells are tetrahedra.
Component Nodes Boundary faces Cells Grid generator
Blades 4,510,177 179, 570 17,687,963 VGRID surface mesh,
(each) AFLR3 volume mesh
Nacelle 971,059 51, 620 2,286,757 VGRID
and tower
Background 4,776,082 10, 488 28,278,639 AFLR3
Total 14,767,495 421, 248 67,084,524
arbitrarily. Any node of one mesh that lies inside solid boundaries defined by another mesh
must be marked in such a way that it does not contribute to the solution. This marking
or mapping procedure must be used to communicate flow information from one mesh to
those it overlaps.
If the meshes are structured, overset methods simplify the grid generation process since
component meshes can be chosen so that it is easy to “wrap” a body-fitted mesh around
each component. Unstructured methodologies do not suffer from the difficulty of fitting a
The composite mesh for the Phase VI rotor consisted of four component meshes: two
blades, the nacelle and tower, and a background mesh. Various statistics for the component
meshes are given in Table 7. The final composite mesh, shown in Figure 42, had over
14.7 million nodes. The blade grids were generated using a combination of Vgrid [73]
and Aflr3 [52]. The former was used to provide a surface grid that was anisotropically
stretched in the spanwise direction. This dramatically reduces the overall node count of
the mesh and is a capability unique to Vgrid. Aflr3 was used to build a mixed-element
volume grid with a boundary layer consisting primarily of prismatic elements. The initial
normal spacing on the blade grids was iteratively refined until y+ < 1 over the entire blade.
Background grid spacing in the vicinity of the rotor was 0.03 m, which is just under 10%
87
(a) Whole domain
(b) Rotor region (c) Nacelle (d) Blade tip cutting through back-
ground
Figure 42: The final composite mesh for the NREL Phase VI turbine.
88
Table 8: Integrated turbine loads. Asterisks in the last column indicate cases that were
impossible to trim. The given pitch is the value that resulted in thrust closest to the target
value.
Wind speed Yaw Turbulence Thrust Torque Root flap Trimmed
(m/s) (deg) model (N) (N-m) bending (N-m) pitch (deg)
Meas. 1664 1313 2471 3.0
10 0 k-ω SST 1821 1808 3055 5.0*
HR-LES 1821 1808 3055 5.0*
Meas. 2278 1104 3040 3.0
15 0 k-ω SST 2660 1670 4165 1.0*
HR-LES 2301 1372 3496 4.0
30 Meas. 2112 1483 2989 3.0
15 HR-LES 2187 1591 3498 2.5
Hole-cutting and grid assembly were accomplished using Suggar++ [64]. Donor inter-
polation within Fun3d was handled by DiRTlib [62]. Due to grid motion, new holes had
to be cut at each time step. This was facilitated by Fun3d’s “DCI-on-the-fly” capability.
8.3 Results
Zero-yaw cases were run at wind speeds of 10 and 15 m/s, using the k-ω SST and HR-LES
(full hybridization) turbulence models for each case. In each case, the rotor was “trimmed”
by manually adjusting the blade pitch until the predicted thrust matched the measured
thrust. This provides a means to correct for any errors in measurement of the blade pitch
in the wind tunnel tests as well as any small differences between the published blade twist
and the twist in the mesh. Thrust, defined here as force directed along the rotor axis that
is averaged over a revolution, is not normally a quantity of interest for wind turbines, but
is less difficult to predict than torque [59]. The nominal tip pitch of 3◦ was used as the
starting point for the trim process. A trimmed solution was not possible in every case.
The manual trimming process requires that thrust be monotonic with pitch, which is not
necessarily the case when the rotor is stalled. In addition, any trim solution that deviated
more than two degrees from the nominal pitch was rejected. The mean loads for each case,
89
8.3.1 10 m/s, 0 degrees yaw
Pressure coefficients at 10 m/s and 0◦ yaw are shown in Figure 43. In the measured pressure,
it is clear that this case has little unsteady flow aside from an area near the leading edge at
r/R = 0.3, as denoted by the bars that indicate the extent of the unsteady measurements.
As a result, the RANS and HR-LES solutions are nearly identical, with no observable
difference in mean pressure. At the three outboard stations, Cp predictions are quite close
to the measured data. At r/R = 0.3, the overall trend is captured, but the leading edge
shedding is not captured, resulting in a lower Cp prediction. The flat Cp profile over the
first half of the chord at r/R = 0.47 is not predicted by either RANS or HR-LES. Due to
the large over-prediction of suction Cp at r/R = 0.47, where the force integration panel area
is equal to 25% of the total planform area, neither RANS nor HR-LES was able to reach
a physically meaningful trimmed solution. A tip pitch of 5◦ was reached before trimming
was abandoned.
The flat Cp profile at r/R = 0.47 indicates separation, and the fact that the separation
occurs at the leading edge suggests that it is a laminar separation, followed by reattachment.
more likely cause of the poor prediction is the low level of freestream turbulence in the
present simulations. The turbulence intensity in the NASA/Ames wind tunnel is roughly
2k/3/U = 0.005 [97], but a value of 7.7 × 10−5 (the Fun3d default) was used here.
p
Wake visualizations for the 10 m/s case with HR-LES are shown in Figure 44. Tip vor-
tices are resolved for 200 degrees of wake age. The disruption of the tip vortices downstream
of the tower is visible in the vortex just behind the tower. The Cp profiles in Figure 43
show fully attached flow in the computational results (suction surface pressure increasing
monotonically from leading to trailing edge), which agrees with the wake visualizations,
Mean pressure coefficients at 15 m/s in Figure 45 show distinct differences between RANS
and HR-LES solutions. At this wind speed, the measured data show that the blade is
90
(a) r/R = 0.30 (b) r/R = 0.47
Figure 43: Mean sectional pressure coefficient at 10 m/s. Bars indicate standard deviation
of pressure at a particular tap, not error. RANS and HR-LES results are coincident.
91
Figure 44: Q-criterion iso-surface (Q = 0.1) at 10 m/s, predicted using HR-LES.
stalled, and the flow is separated everywhere but the tip. At all stations except the tip,
HR-LES yields a more accurate prediction of the mean pressure. At the three mid-span
stations, HR-LES predicts the experimentally observed stalled flow where RANS does not.
Neither turbulence model correctly predicts the experimental pressure at r/R = 0.3. HR-
LES captures the basic character of the stalled and separated leading edge, but reattaches
near 20% chord rather than 55%. This may again be due to a freestream turbulence intensity
substantially lower than the experimental value. For this case, RANS was unable to trim
due to consistent over-prediction of suction at r/R = 0.3, 0.47, and 0.8, and trimming ended
at 1◦ tip pitch. HR-LES, on the other hand, was able to predict the target thrust within
Wake visualizations for the 15 m/s case are shown in Figures 46 and 47. Clearly, HR-
LES is able to predict the massive separation from the suction surfaces, as exhibited by the
strong vortices visible in Figure 47, shedding from mid-span and persisting for over 90◦ of
wake age. In the RANS solution, some vorticity is shed into the wake from mid-span, but
92
(a) r/R = 0.30 (b) r/R = 0.47
93
Figure 46: Q-criterion iso-surface (Q = 0.1) at 15 m/s, predicted using RANS.
it dissipates much faster than with HR-LES. Both models capture well-defined tip and root
vortices, but those vortices are smoother in the RANS solution. In contrast, in the HR-LES
solution, the tip vortices interact with the separated wake as they convect downstream,
One case was run at 15 m/s and 30◦ yaw. Since yawed cases are always unsteady, no
matter the wind speed, only the HR-LES turbulence model was used here. Mean pressure
coefficients are shown in Figure 48. Outboard pressure predictions are observed to be very
accurate when correlated with the measured data. At r/R = 0.3 and r/R = 0.47, the flow is
considerably less steady, as evidenced by the size of the bars on the measured data in Figure
48. The predicted pressure is within a standard deviation of the measured mean at most
pressure tap locations. With such a chaotic wake, some disagreement is to be expected in
mean pressure since the computational results have been averaged over a single revolution
94
Figure 47: Q-criterion iso-surface at 15 m/s, predicted using HR-LES.
Integrated hub loads vs. azimuth for the 15 m/s, 30◦ yaw case are shown in Figure
49. The current methodology predicts thrust and torque within 4% and 7%, respectively,
of the targets. The unsteadiness observed in the computed thrust between 0◦ and 180◦
azimuth is comparable to the wide spread in the measured data, which was gathered over
36 revolutions. In addition to a larger apparent angle of attack, this region is in the shadow
of the nacelle, and in the experimental data, the hub instrumentation box and boom. The
same can be seen in root flap bending moment, which is also derived from sectional normal
force, but with greater weight given to the tip. The maximum torque is somewhat over-
predicted, which could be due to the aforementioned transition effect [41]. The maximum
Sectional normal force and pitching moment vs azimuth are shown in Figures 50 and 51.
Similar to the integrated loads, the wide spread in the measured data in the downwind half
of the rotor is well-predicted. At very high yaw angles, the instrumentation box and camera
95
(a) r/R = 0.30 (b) r/R = 0.47
96
(a) Thrust
(c) Torque
Figure 49: Integrated hub loads (pressure component only) vs. azimuth at 15 m/s, 30◦ yaw.
97
boom mounted in front of the hub (see Figure 41) have a significant effect on blade pressure
measurements in the 0 ≤ ψ < 120◦ region [35] (ψ is zero for a blade at the 12 o’clock position
and increases for counter-clockwise rotation when viewed from upstream). However, at 30◦
yaw, only r/R < 0.22 (approximately) is in the box’s shadow. The boom’s shadow extends
to about r/R < 0.34. Therefore, only the first pressure tap station, at r/R = 0.3 should be
affected by the omission of the box and boom in the current simulations. Any unsteadiness
at the outer four locations is due to the changing local angle of attack.
expected in a yawed case. A larger extent of periodic shedding can be observed for the
blade that is in the 6 o’clock position than for the blade at 12 o’clock. The 90◦ ≤ ψ < 270◦
region experiences a higher incident velocity than the other half of the rotor disc. Therefore,
Using an incompressible solution algorithm was essential to the results presented here.
Figure 53 compares the wake presented earlier for the 15 m/s, 0◦ yaw case, to the same
case computed with a compressible method [105] (without low-Mach preconditioning, but
using HR-LES). The compressible algorithm is much more dissipative at this Mach number,
leading to less well-defined tip and root vortices. The same is true of the mid-span separation
equation need not be solved, so the mean flow Jacobian consists of 4 × 4 rather than 5 × 5
blocks.
8.4 Performance
All flow solutions presented here were run on the Intel 64 Cluster “Abe” at the National
Center for Supercomputing Applications (NCSA). Each cluster node had one 2.33 GHz
quad-core CPU. As in the HART simulations, 400 cores were used for the flow solution and
a single core for hole-cutting. At 720 steps per revolution, one revolution took 25–30 hours
of wall time. Trimming required 3–4 revolutions. Fifteen percent of the total solution time
98
(a) r/R = 0.30 (b) r/R = 0.47
Figure 50: Sectional normal force coefficient, normalized by tip dynamic pressure, vs. az-
imuth at 15 m/s, 30◦ yaw.
99
(a) r/R = 0.30 (b) r/R = 0.47
Figure 51: Sectional pitching moment coefficient, normalized by tip dynamic pressure, vs.
azimuth at 15 m/s, 30◦ yaw.
100
Figure 52: Q-criterion iso-surface (Q = 0.1) at 15 m/s and 30◦ yaw, predicted using HR-
LES.
Figure 53: Q-criterion iso-surface at 15 m/s and 0◦ yaw, computed using a compressible
algorithm [105].
101
CHAPTER IX
CONCLUSIONS
A range of different CFD techniques of varying accuracy and computational expense has
been developed for simulating horizontal axis wind turbines. Improved unsteady actua-
tor blade methods for efficiently simulating wind turbine wakes without fully resolving the
blades were developed. Possible applications for these methods include turbine-to-turbine
interaction studies. A hybrid RANS/LES turbulence model was modified for unstructured
cross-flow and flatback airfoil. A tight CFD/CSD coupling methodology for very high
fidelity overset simulations of wind turbines was demonstrated. Unstructured overset simu-
lations of a wind turbine using a hybrid RANS/LES turbulence model were performed and
correlated with an experimental dataset. All overset turbine simulations were conducted in
a time-accurate fashion in an inertial reference frame and included the turbine tower and
• An improved nearest neighbor search algorithm based on kd-trees was shown to de-
crease the CPU time for source-to-node association by as much as 85% when there is
a large number of mesh nodes per CPU. This can result in a 20% reduction in total
• It was also shown that using the actual non-linear twist distribution rather than a
linear approximation of the twist was necessary for accurate actuator wake predictions.
ment at 7 m/s were compared against high-fidelity overset simulations that model the
real blade geometry. It was shown that the actuator blade method predicts tip and
root vortex trajectories nearly identical to the overset method in the near wake with
102
a 38% reduction in CPU time per degree of freedom compared to the overset solution.
• The HRLES model yields improved predictions of drag and shedding frequency for
• These models correctly predict the unsteady, 3-D, and chaotic nature of bluff-body
wakes.
• The HRLES model developed here is sensitive to mesh cells with a very high aspect
ratio. Non-physical solutions were observed for meshes containing tetrahedra with
aspect ratios greater than 8 close to the wall but outside the boundary layer. Most
unstructured grid generators automatically generate low aspect ratio cells outside the
boundary layer (they are unavoidable in the boundary layer), but trouble can arise
• Just as tight coupling can be used to simulate maneuvering flight in helicopters [66, 2],
the process can be applied to with turbine simulations that include yaw adjustments
and momentary wind gusts, neither of which have a quasi-periodic loading as is re-
• Though the simulation used to demonstrate tight CFD/CSD coupling was carried out
using a helicopter rotor for which an experimental dataset was available for verifica-
tion, the principles are similar for a wind turbine rotor (see future work below).
• Should an experimental dataset for a wind turbine rotor aeroelastic effects more typ-
ical of a utility-scale turbine become publicly available in the future, this coupling
methodology can then be demonstrated for that rotor without any modifications to
the algorithm.
103
• At 10 m/s wind speed, RANS and HRLES methods were found to yield nearly identical
solutions as the boundary layer remains attached over most of the rotor.
• Poor prediction of the peak suction at r/R = 0.47 in the 10 m/s case caused both
RANS and HRLES to over-predict rotor thrust and torque, precluding a trimmed
• At 15 m/s, where the rotor is almost fully stalled, HRLES was able to predict thrust
by 51%, HRLES predicted torque within 16% of the measured value. The improved
loads predictions were due to the ability of HRLES to capture the highly unsteady
• HRLES was also able to accurately capture blade loads in yawed flow with excellent
mean Cp correlations across the span. Wake asymmetry was accurately captured as
well as increased unsteadiness in sectional loads in the downwind half of the rotor
disc.
A shortcoming of the actuator blade methodology is that the source terms that drive
the CFD solution are based on extremely simple linear, quasi-steady aerodynamics.
Much more accurate approximations for unsteady aerodynamics exist in the form of
finite state aerodynamics algorithms, such as those of Peters [72]. These methods are
still very efficient compared to full overset CFD. Since they already exist in many com-
prehensive codes like Dymore, coupling with a comprehensive code could introduce
these unsteady effects with relatively little effort in implementation. The CFD code
would provide local angles of attack to the comprehensive code, which would then
provide instantaneous lift and drag, which can then be used to compute the source
104
terms. Coupling of a steady-state CFD actuator disc with a comprehensive code was
to time-accurate simulations.
• Non-rectangular planforms
Presently, the actuator blade method does not allow for non-rectangular planforms,
but modern wind turbine blades are typically tapered. Furthermore, novel designs,
such as those with sweep twist adaptive rotors (STAR) [7] may have blades that are
curved and swept as well as tapered. Due to the rectangular assumption, sources
will not be located in the correct locations to properly approximate the real rotor.
that the low Reynolds number was responsible for many of the disagreements with
• The HRLES model should be extended to utilize the dynamic closure coefficients
specified in the LDKM model. This would make the model more generally applicable
• The characteristic grid length used in the LES destruction term should be revisited.
Spalart [100] notes that on unstructured grids, (dual) cell diameter would be a more
appropriate estimate of mesh size than volume1/3 . Isotropic LES models, like the one
used in the hybrid model here, have no preferred eddy orientation, so characteristic
length should ideally be the largest distance between a node and its neighbors in any
direction.
105
• The flatback airfoil simulations with tunnel walls should be revisited once the wall
porosity is known. Given the large tunnel blockage, modeling the solid upper and
lower walls can be expected to improve correlations with measured data but only if
Possible areas for future development of overset wind turbine simulation techniques
include:
Since many wind turbines are designed with laminar flow airfoils, and do indeed have
could range from simple methods like Michel’s criterion [57] to high-fidelity meth-
transport equations. Of course, the computational cost of the latter approach would
laminar solver.
Algorithms for computing the trimmed state of a rigid rotor would be substantially
reduce this cost would dramatically improve performance of overset simulations. Im-
mesh partitioning strategy for rotor simulations could trade some solver inefficiency
have already been explored by Noack [64], but their effect on solver efficiency has not
106
been determined.
As more experimental and field datasets become available, they should be used to
provide further correlations for evaluating the tight coupling methods developed here.
Predictions of structural responses to gusts and the vertical shear profile would be
wind turbines, such as flexible towers, non-uniform freestream, and non-uniform blade
Both the approximate and high-fidelity prediction methods developed here could be
used to explore the effects of wind shear in the atmospheric boundary layer on rotor
The basic physical processes in the development of wind turbine wakes should be
studied with an emphasis on how turbine wakes differ from helicopter wakes. Though
there are obvious similarities between helicopter and wind turbine rotors, there should
will require the application of higher-order algorithms and/or finer grid resolution (via
either uniform or adaptive mesh refinement) to resolve tip vortices at long wake age.
107
REFERENCES
[1] “20% Wind Energy by 2030,” Tech. Rep. DOE/GO-102008-2567, U.S. Department of
Energy, July 2008.
[2] Abhishek, A., Datta, A., and Chopra, I., “Comprehensive Analysis, Prediction,
and Validation of UH-60A Blade Loads in Unsteady Maneuvering Flight,” in Ameri-
can Helicopter Society 63rd Annual Forum, (Virginia Beach, VA), May 1-3, 2007.
[3] Abras, J. N., Enhancement of Aeroelastic Rotor Airload Prediction Methods. PhD
thesis, Georgia Institute of Technology, 2009.
[4] Abras, J. N., Lynch, C. E., and Smith, M. J., “Advances in Rotorcraft Simula-
tions with Unstructured CFD,” in American Helicopter Society 63rd Annual Forum,
(Virginia Beach, VA), May 1-3, 2007.
[5] Ahmad, J. and Duque, E., “Helicopter Rotor Blade Computation in Unsteady Flows
Using Moving Overset Grids,” Journal of Aircraft, vol. 33, no. 1, pp. 54–60, 1996.
[6] Anderson, W., Rausch, R., and Bonhaus, D., “Implicit/Multigrid Algorithms for
Incompressible Turbulent Flows on Unstructured Grids,” Journal of Computational
Physics, vol. 128, no. 2, pp. 391–408, 1996.
[7] Ashwill, T., Kanaby, G., Jackson, K., and Zuteck, M., “Development of
the Sweep-Twist Adaptive Rotor (STAR) Blade,” in 48th AIAA Aerospace Sciences
Meeting and Exhibit, (Orlando, FL), AIAA-2010-1582, Jan. 4-7, 2010.
[8] Bauchau, O. A. and Ahmad, J. U., “Advanced CFD and CSD Methods for Multi-
disciplinary Applications in Rotorcraft Problems,” in 6th AIAA, NASA, and ISSMO
Symposium on Multidisciplinary Analysis and Optimization, (Bellevue, WA), AIAA-
1996-4151, Sept. 4-6, 1996.
[10] Bauchau, O. A., Bottasso, C. L., and Nikishkov, Y. G., “Modeling Rotorcraft
Dynamics with Finite Element Multibody Procedures,” Mathematical and Computer
Modeling, vol. 33, no. 10-11, pp. 1113–1137, 2001.
[11] Bauchau, O.A., “Dymore User’s Manual,” tech. rep., Georgia Institute of Technol-
ogy, June 2010.
[12] Baurle, R. A., Tam, C. J., Edwards, J. R., and Hassan, H. A., “Hybrid Sim-
ulation Approach for Cavity Flows: Blending Algorithm, and Boundary Treatment
Issues,” AIAA Journal, vol. 41, no. 8, pp. 1463–1480, 2003.
[13] Bentley, J. L., “Multidimensional Binary Search Trees Used for Associate Search-
ing,” Communications of the ACM, vol. 18, no. 9, pp. 509–517, 1975.
108
[14] Berg, D. and Zayas, J., “Aerodynamic and Aeroacoustic Properties of Flatback
Airfoils,” in 46th AIAA Aerospace Sciences Meeting and Exhibit, (Reno, NV), AIAA-
2008-1455, Jan. 5-8, 2008.
[15] Biedron, R., Vatsa, V., and Atkins, H., “Simulation of Unsteady Flows Using an
Unstructured Navier-Stokes Solver on Moving and Stationary Grids,” in 23rd AIAA
Applied Aerodynamics Conference, (Toronto, Canada), AIAA-2005-5093, June 6-9,
2005.
[16] Biedron, R. T. and Lee-Rausch, E. M., “Rotor Airloads Prediction Using Un-
structured Meshes and Loose CFD/CSD Coupling,” in 26th AIAA Applied Aerody-
namics Conference, (Honolulu, HI), AIAA-2008-7341, Aug. 18-21, 2008.
[17] Bonhaus, D., “An Upwind Multigrid Method For Solving Viscous Flows On Un-
structured Triangular Meshes,” Master’s thesis, George Washington University, 1993.
[18] Bridgeman, J. O., Strawn, R. C., Caradonna, F. X., and Chen, C. S., “Ad-
vanced Rotor Computations with a Corrected Potential Method,” in American Heli-
copter Society 45th Annual Forum Proceedings, (Boston, MA), May 22-24 1989.
[19] Buning, P., Parks, S., Chan, W., and Renze, K., “Application of the Chimera
Overlapped Grid Scheme to Simulation of Space Shuttle Ascent Flows,” in Proceedings
of the Fourth International Symposium on Computational Fluid Dynamics, vol. 1,
(Davis, CA), pp. 132–137, Sept. 1991.
[20] Cantaloube, B. and Huberson, S., “A New Approach Using Vortex Point Method
for Prediction of Rotor Performance in Hover and Forward Flight,” Vertica, vol. 10,
no. 2, pp. 187–200, 1986.
[22] Chao, D. D. and van Dam, C. P., “RaNS Analysis of an Inboard Flatback Modifi-
cation to the NREL Phase VI Rotor,” in 44th Aerospace Sciences Meeting and Exhibit,
(Reno, NV), AIAA-2006-0195, Jan. 9-12, 2006.
[23] Chorin, A., “A Numerical Method for Solving Incompressible Viscous Flow Prob-
lems,” Journal of Computational Physics, vol. 2, no. 1, pp. 12–26, 1967.
[24] Cockle, R., “Wind Whips Up Health Fears,” The Oregonian, Aug. 10, 2008.
[25] Coton, F. N., Wang, T., and Galbraith, R. A. M., “An Examination of Key
Aerodynamic Modeling Issues Raised by the NREL Blind Comparison,” in 21st ASME
Wind Energy Symposium and 40th AIAA Aerospace Sciences Meeting and Exhibit,
(Reno, NV), AIAA-2002-0038, Jan. 14-17, 2002.
[26] Datta, A., Sitaraman, J., Baeder, J. D., and Chopra, I., “Analysis Refinements
for Prediction of Rotor Vibratory Loads in High Speed Forward Flight,” in American
Helicopter Society 60th Annual Forum Proceedings, (Baltimore, MD), June 7-10, 2004.
109
[27] Davidson, P. A., Turbulence: An Introduction for Scientists and Engineers. New
York: Oxford University Press, 2004.
[28] Duque, E. P. N., Burklund, M. D., and Johnson, W., “Navier-Stokes and Com-
prehensive Analysis Performance Predictions of the NREL Phase VI Experiment,”
Journal of Solar Energy Engineering, vol. 125, pp. 457–467, November 2003.
[29] Duque, E. P. N., van Dam, C. P., and Hughes, S. C., “Navier-Stokes Simulations
of the Combined Experiment Phase II Rotor,” in ASME Wind Energy Symposium,
(Reno, NV), AIAA-1999-0037, Jan. 11-14, 1999.
[30] Duque, E. et al., “Revolutionary Physics-Based Design Tools for Quiet Heli-
copters,” in 44th AIAA Aerospace Sciences Meeting and Exhibit, (Reno, NV), AIAA-
2006-1068, Jan. 9-12, 2006.
[32] Friedman, J. H., Bentley, J. L., Finkeyl, R. A., “An Algorithm for Finding
Best Matches in Logarithmic Expected Time,” Tech. Rep. STAN-CS-75-482, Stanford
CS Rep., Feb. 1975.
[33] Gaither, J. A., Marcum, D. L., and Mitchell, B., “Solidmesh: A Solid Model-
ing Approach to Unstructured Grid Generation,” in 7th International Conference on
Numerical Grid Generation in Computational Field Simulations, (Whistler, Canada),
Sept. 25-28, 2000.
[34] Grosveld, F. W., “Prediction of Broadband Noise from Horizontal Axis Wind
Turbines,” Journal of Propulsion and Power, vol. 1, no. 4, pp. 292–299, 1985.
[35] Hand, M. M. et al., “Unsteady Aerodynamics Experiment Phase VI: Wind Tun-
nel Test Configurations and Available Data Campaigns,” Tech. Rep. NREL/TP-500-
29955, National Renewable Energy Laboratory, Golden, CO, Dec. 2001.
[36] Huyer, S. A., Simms, D., Robinson, M. C., “Unsteady Aerodynamics Associated
with a Horizontal-Axis Wind Turbine,” AIAA Journal, vol. 34, no. 7, pp. 1410–1419,
1996.
[38] Ivanell, S., Sørensen, J. N., Mikkelsen, R., and Henningson, D., “Analysis
of Numerically Generated Wake Structures,” Wind Energy, vol. 12, no. 1, pp. 63–80,
2009.
[39] Kim, W.-W. and Menon, S., “An Unsteady Incompressible Navier-Stokes Solver
for Large Eddy Simulation of Turbulent Flows,” International Journal for Numerical
Methods in Fluids, vol. 31, no. 6, pp. 983–1017, 1999.
[40] Kravchenko, A. G. and Moin, P., “Numerical Studies of Flow Over a Circular
Cylinder at ReD = 3900,” Physics of Fluids, vol. 12, no. 2, pp. 403–417, 2000.
110
[41] Langtry, R., Gola, J., and Menter, F., “Predicting 2D Airfoil and 3D Wind
Turbine Rotor Performance Using a Transition Model for General CFD Codes,” in
44th AIAA Aerospace Sciences Meeting and Exhibit, (Reno, NV), AIAA-2005-0395,
Jan. 9-12, 2006.
[42] Le Chuiton, F., “Actuator Disc Modeling for Helicopter Rotors,” Aerospace Science
and Technology, vol. 8, no. 4, pp. 285–297, 2004.
[44] Lee, C. S., Saberi, H., and Ormiston, R. A., “Aerodynamic and Numerical
Issues for Coupling CFD into Comprehensive Rotor Analysis,” in American Helicopter
Society 53rd Annual Forum Proceedings, (Virginia Beach, VA), April 29–May 1, 1997.
[45] Leishman, J. G., “Challenges in Modeling the Unsteady Aerodynamics of Wind Tur-
bines,” in 21st ASME Wind Energy Symposium and 40th AIAA Aerospace Sciences
Meeting and Exhibit, (Reno, NV), AIAA-2002-0037, Jan. 14-17, 2002.
[46] Lim, J. W., Nygaard, T. A., Strawn, R., and Potsdam, M., “BVI Airloads Pre-
diction Using CFD/CSD Loose Coupling,” in AHS San Francisco Bay Area Chapter
Specialists’ Conference on Vertical Lift Aircraft Design, (San Francisco, CA), Jan.
18-20, 2006.
[47] Lim, J. and Strawn, R., “Prediction of HART II Rotor BVI Loading and Wake
System Using CFD/CSD Loose Coupling,” in 45th AIAA Aerospace Sciences Meeting
and Exhibit, (Reno, NV), AIAA-2007-1281, Jan. 8-11, 2007.
[48] Lourenco, L.M. and Shih, C., “Characteristics of the Plane Turbulent Near Wake
of a Circular Cylinder.” Digitized from Figures 13, 15, and 22 in Reference [40].
[49] Madsen, H. and Johansen. J. and Sørensen, N. and Larsen, G. and Hansen,
M., “Simulation of Low Frequency Noise from a Downwind Wind Turbine Rotor,” in
45th AIAA Aerospace Sciences Meeting and Exhibit, (Reno, NV), AIAA-2007-0623,
Jan. 8-11, 2007.
[50] Makinen, S., Hill, M., Gandhi, F., Long, L. N., Vasilescu, R., and Sankar,
L., “A Study of the HART-I Rotor with Loose Computational Fluid/Structural Dy-
namic Coupling,” in American Helicopter Society 62nd Annual Forum Proceedings,
(Phoenix, AZ), May 9-11, 2006.
[51] Marcum, D. L., “Unstructured Grid Generation Using Automatic Point Insertion
and Local Reconnection,” in The Handbook of Grid Generation (Thompson, J. F.,
Soni, B., and Weatherill, N. P., eds.), ch. 18, CRC Press, 1998.
[53] Meakin, R., “Moving Body Overset Grid Methods for Complete Aircraft Tiltrotor
Simulations,” in 11th AIAA Computational Fluid Dynamics Conference, (Orlando,
FL), AIAA-1993-3350, July 6-9, 1993.
111
[54] Menter, F. R., “Improved Two-Equation k-ω Turbulence Models for Aerodynamic
Flows,” Tech. Rep. NASA TM 103975, NASA, Oct. 1992.
[55] Menter, F., “Two-Equation Eddy-Viscosity Turbulence Models for Engineering Ap-
plications,” AIAA Journal, vol. 32, no. 8, pp. 598–605, 1994.
[56] Menter, F. et al., “Correlation-Based Transition Model Using Local Variables—
Part I: Model Formulation,” Journal of Turbomachinery, vol. 128, no. 3, pp. 413–422,
2006.
[57] Michel, R., “Etude de la Transition sur les Profiles d’Aile; Establissement d’un
Critère de Determination de Point de Transition et Calcul de la Trainée de Profile
Incompressible,” Tech. Rep. 1/1578A, ONERA, 1951.
[58] Mikkelsen, R., Actuator Disc Methods Applied to Wind Turbines. PhD thesis,
Technical University of Denmark, 2003.
[59] Montaudouin, J., Heo, S., Smith, M., and Bauchau, O., “Aerodynamic and
Aeroelastic Analysis of Rotors at High Advance Ratio,” in 36th European Rotorcraft
Forum, (Paris, France), Sept. 7-9, 2010.
[60] Moriarty, P.J., “AeroDyn Theory Manual,” Tech. Rep. NREL/EL-500-36881, Na-
tional Renewable Energy Laboratory, Dec. 2005.
[61] Narducci, R., “Comparison of Blade Tip Vortex Calculations to Wind Tunnel Mea-
surements,” in Proceedings of the Army Research Office Rotorcraft Wake Prediction
Basic Research Workshop, (Georgia Institute of Technology, Atlanta, GA), Mar. 16-
17, 2009.
[62] Noack, R., “DiRTlib: A Library to Add an Overset Capability to Your Flow
Solver,” in 17th AIAA Computational Fluid Dynamics Conference, (Toronto, On-
tario, Canada), AIAA-2005-5116, June 6-9, 2005.
[63] Noack, R., “SUGGAR: A General Capability for Moving Body Overset Grid Assem-
bly,” in 17th AIAA Computational Fluid Dynamics Conference, (Toronto, Ontario,
Canada), AIAA-2005-5117, June 6-9, 2005.
[64] Noack, R., Boger, D., Kunz, R., and Carrica, P., “Suggar++: An Improved
General Overset Grid Assembly Capability,” in 19th AIAA Computational Fluid Dy-
namics Conference, (Austin, TX), AIAA-2009-3992, June 22-25, 2009.
[65] Norberg, C. Unpublished data digitized from Figure 11 in Reference [40].
[66] Nygaard, T., Saberi, H., Ormiston, R., Strawn, R., and Potsdam, M., “CFD
and CSD Coupling Algorithms and Fluid Structure Interface for Rotorcraft Aerome-
chanics in Steady and Transient Flight Conditions,” in American Helicopter Society
62nd Annual Forum Proceedings, (Phoenix, AZ), May 9-11, 2006.
[67] O’Brien, D. M., Analysis of Computational Modeling Techniques for Complete Ro-
torcraft Configurations. PhD thesis, Georgia Institute of Technology, 2006.
[68] O’Brien, D. M. and Smith, M. J., “Analysis of Rotor-Fuselage Interactions Us-
ing Various Rotor Models,” in AIAA 43rd Aerospace Sciences Meeting, (Reno, NV),
AIAA-2005-0468, Jan. 10-13, 2005.
112
[69] Obrien, D. M. and Smith, M. J., “Understanding the Physical Implications of
Approximate Rotor Methods Using an Unstructured CFD Method,” in 31st European
Rotorcraft Forum, (Florence, Italy), Sept. 13-15, 2005.
[70] Ong, L. and Wallace, J., “The Velocity Field of the Turbulent Very Near Wake
of a Circular Cylinder,” Experimental Fluids, vol. 20, no. 6, pp. 441–453, 1996.
[71] Paquette, J. A. and Veers, P. A., “Increased Strength in Wind Turbine Blades
Through Innovative Structural Design,” in Proceedings of the European Wind Energy
Conference, (Milan, Italy), May 7-10, 2007.
[72] Peters, D. A., Karunamoorthy, S., and Cao, Wen-Ming, “Finite State In-
duced Flow Models Part I: Two-Dimensional Thin Airfoil,” Journal of Aircraft,
vol. 32, no. 2, pp. 313–321, 1995.
[73] Pirzadeh, S., “Three-Dimensional Unstructured Viscous Grids by the Advancing
Front Method,” AIAA Journal, vol. 34, no. 1, pp. 43–49, 1996.
[74] Potsdam, M. and Mavriplis, D., “Unstructured Mesh CFD Aerodynamic Analysis
of the NREL Phase VI Rotor,” in 47th AIAA Aerospace Sciences Meeting, (Orlando,
FL), AIAA-2009-1221, Jan. 5-8, 2009.
[75] Potsdam, M., Smith, M., and Renaud, T., “Unsteady Computations of Rotor-
Fuselage Interaction,” in 35th European Rotorcraft Forum, (Hamburg, Germany),
Sept. 22-25, 2009.
[76] Potsdam, M. A., Venkateswaran, S., and Pandya, S., “Unsteady Low Mach
Preconditioning with Application to Rotorcraft Flows,” in 18th AIAA Computational
Fluid Dynamics Conference, (Miami, FL), AIAA-2007-4473, June 25-28, 2007.
[77] Potsdam, M., Yeo, H., and Johnson, W., “Rotor Airloads Prediction Using Loose
Aerodynamic/Structural Coupling,” in American Helicopter Society 60th Annual Fo-
rum Proceedings, (Baltimore, MD), June 7-10, 2004.
[78] Renaud, T., D. M. O’Brien, J., Smith, M. J., and Potsdam, M., “Evalua-
tion of Isolated Fuselage and Rotor-Fuselage Interaction Using CFD,” Journal of the
American Helicopter Society, vol. 53, no. 1, pp. 3–17, 2008.
[79] Sanchez-Rocha, M., Kirtas, M., and Menon, S., “Zonal Hybrid RANS-LES
Method for Static and Oscillating Airfoils and Wings,” in 44th AIAA Aerospace Sci-
ences Meeting and Exhibit, (Reno, NV), AIAA-2006-1256, Jan. 9-12, 2006.
[80] Sanchez-Rocha, M., Wall-Models for Large Eddy Simulation Based on a Generic
Additive-Filter Formulation. PhD thesis, Georgia Institute of Technology, 2008.
[81] Schlichting, H. and Gersten, K., Boundary Layer Theory. Springer, 8 ed., 2000.
[82] Schreck, S., “Two-Dimensional Unsteady Aerodynamic Processes in Low Frequency
Shedding under Rotational Augmentation,” in 47th AIAA Aerospace Sciences Meeting
and Exhibit, (Orlando, FL), AIAA-2009-1222, Jan. 5-8, 2009.
[83] Schreck, S., “Low Frequency Shedding Prompted by Three Dimensionality Under
Rotational Augmentation,” in 48th AIAA Aerospace Sciences Meeting and Exhibit,
(Orlando, FL), AIAA-2010-0640, Jan. 4-7, 2010.
113
[84] Schreck, S. and Robinson, M., “Blade Three-Dimensional Dynamic Stall Re-
sponse to Wind Turbine Operating Conditions,” Journal of Solar Energy Engineering,
vol. 127, no. 4, pp. 488–495, 2005.
[85] Schreiber, O., Aerodynamic Interactions Between Bodies in Relative Motion. PhD
thesis, Georgia Institute of Technology, 1990.
[86] Schweikhard, R., “Actuator Disk for Helicopter Rotors in an Unstructured Flow
Solver,” Journal of the American Helicopter Society, vol. 52, no. 1, pp. 58–68, 2007.
[87] Sezer-Udol, N. and Long, L. N., “3-D Time-Accurate CFD Simulations of Wind
Turbine Rotor Flow Fields,” in 44th AIAA Aerospace Sciences Meeting and Exhibit,
(Reno, NV), AIAA-2006-0394, Jan. 9-12, 2006.
[88] Shelton, A. B., Braman, K., Smith, M. J., and Menon, S., “Improved Hy-
brid RANS-LES Turbulence Modeling for Rotorcraft,” in American Helicopter Society
62nd Annual Forum Proceedings, (Phoenix, AZ), May 9-11, 2006.
[89] Shelton, A., A Multi-resolution Discontinuous Galerkin Method for Unsteady Com-
pressible Flows. PhD thesis, Georgia Institute of Technology, 2008.
[90] Simms, D., Schreck, S., Hand, M., and Fingersh, L. J., “NREL Unsteady Aero-
dynamics Experiment in the NASA-Ames Wind Tunnel: A Comparison of Predic-
tions to Measurements,” Tech. Rep. NREL/TP-500-29494, National Renewable En-
ergy Laboratory, Golden, CO, June 2001.
[91] Sitaraman, J., Katz, A., Jayaraman, B., Wissink, A., and Sankaran, V.,
“Evaluation of a Multi-Solver Paradigm for CFD using Unstructured and Adaptive
Cartesian Grids,” in 46th AIAA Aerospace Sciences Meeting and Exhibit, (Reno, NV),
AIAA-2008-0660, Jan. 7-10, 2008.
[92] Smith, M. J, A Fourth Order Euler/Navier-Stokes Prediction Method for the Aero-
dynamics and Aeroelasticity of Hovering Rotor Blades. PhD thesis, Georgia Institute
of Technology, 1994.
[93] Smith, M. J. and Potsdam, M. and Wong, T.-C. and Baeder, J. D. and
Phanse, S., “Evaluation of CFD to Determine Two-Dimensional Airfoil Characteris-
tics for Rotorcraft Applications,” Journal of the American Helicopter Society, vol. 51,
no. 1, pp. 70–79, 2006.
[94] Smith, M. J., Koukol, B. C. G., Quackenbush, T., and Wachspress, D.,
“Reverse- and Cross-Flow Aerodynamics for High-Advance Ratio Flight,” in 35th
European Rotorcraft Forum, (Hamburg, Germany), Sept. 22-25, 2009.
[95] Son, J. and Hanratty, T. J., “Velocity Gradients at the Wall for Flow Around a
Cylinder at Reynolds Numbers 5 × 103 to 105 ,” Journal of Fluid Mechanics, vol. 35,
no. 2, pp. 353–368, 1969.
[96] Sørensen, J. N. and Shen, W. Z., “Numerical Modeling of Wind Turbine Wakes,”
Journal of Fluids Engineering, vol. 124, no. 2, pp. 393–399, 2002.
[97] Sørensen, N. N., “CFD Modelling of Laminar-turbulent Transition for Airfoils and
Rotors Using the γ-Re
f θ Model,” Wind Energy, vol. 12, no. 8, pp. 715–733, 2009.
114
[98] Spalart, P. and Allmaras, S., “A One-Equation Turbulence Model for Aerody-
namic Flows,” in 30th AIAA Aerospace Sciences Meeting and Exhibit, (Reno, NV),
AIAA-1992-0439, Jan. 6-9, 1992.
[101] Speziale, C. G., “Turbulence Modeling for Time-Dependent RANS and VLES: A
Review,” AIAA Journal, vol. 36, no. 2, pp. 173–184, 1998.
[102] Standish, K. J. and van Dam, C. P., “Analysis of Blunt Trailing Edge Airfoils,”
in 41st Aerospace Sciences Meeting and Exhibit, (Reno, NV), AIAA-2003-0353, Jan.
6-9, 2003.
[103] Stone, C. P., Tebo, S. M., and Duque, E. P. N., “Computational Fluid Dynamics
of Flatback Airfoils for Wind Turbine Applications,” in 44th AIAA Aerospace Sciences
Meeting and Exhibit, (Reno, NV), AIAA-2006-0194, Jan. 9-12, 2006.
[104] Stone, C. and Barone, M. and Lynch, C. E. and Smith, M. J., “A Compu-
tational Study of the Aerodynamics and Aeroacoustics of a Flatback Airfoil Using
Hybrid RANS-LES,” in 47th AIAA Aerospace Sciences Meeting and Exhibit, (Or-
lando, FL), AIAA-2009-0273, Jan. 5-8, 2009.
[105] Stone, C., Lynch, C.E., and Smith, M.J., “Hybrid RANS/LES Simulations of
a Horizontal Axis Wind Turbine,” in 48th AIAA Aerospace Sciences Meeting and
Exhibit, (Orlando, FL), AIAA-2010-0459, Jan. 4-7, 2010.
[106] Strelets, M., “Detached Eddy Simulation of Massively Separated Flows,” in 39th
AIAA Aerospace Sciences Meeting and Exhibit, (Reno, NV), AIAA-2001-0879, Jan.
8-11, 2001.
[107] Szydlowski, J. and Costes, M., “Simulation of Flow Around a NACA 0015 Airfoil
for Static and Dynamic Stall Configurations using RANS and DES,” in AHS Interna-
tional 4th Decennial Specialists’ Conference on Aeromechanics, San Francisco, CA,
Jan. 21-23, 2004.
[108] Thepvongs, S., Cook, J. R., Cesnik, C. E. S., and Smith, M. J., “Computational
Aeroelasticity of Rotating Wings with Deformable Airfoils,” in American Helicopter
Society 65th Annual Forum Proceedings, (Grapevine, TX), May 27-29, 2009.
[109] Travin, A., Shur, M., Strelets, M., and Spalart, P. R., “Detached-Eddy
Simulation Past a Circular Cylinder,” Flow, Turbulence and Combustion, vol. 63,
no. 1-4, pp. 293–313, 1999.
[110] Tung, C., Caradonna, F. X., Boxwell, D. A., and Johnson, W. R., “The
Prediction of Transonic Flows on Advancing Rotors,” in American Helicopter Society
40th Annual Forum Proceedings, (Arlington, VA), May 16-18, 1984.
115
[111] Vatsa, V. N. and Carpenter, M. H., “Higher Order Temporal Schemes with Error
Controllers for Unsteady Navier-Stokes Equations,” in 17th AIAA Computational
Fluid Dynamics Conference, (Toronto, Ontario, Canada), AIAA-2005-5245, June 6-9,
2005.
[113] Wilcox, D.C., Turbulence Modeling for CFD. CDW Industries, 2nd ed., 1998.
[114] Wissink, A., Sitaraman, J., Sankaran, V., Mavriplis, D., and Pulliam, T.,
“Multi-Code Python-Based Infrastructure for Overset CFD with Adaptive Cartesian
Grids,” in 46th AIAA Aerospace Sciences Meeting and Exhibit, (Reno, NV), AIAA-
2008-0927, Jan. 7-10, 2008.
[115] Xu, G. and Sankar, L., “Development of Engineering Aerodynamics Models Using
a Viscous Flow Methodology on the NREL Phase VI Rotor,” Wind Energy, vol. 5,
no. 2-3, pp. 171–183, 2002.
[116] Yu, Y. H., Tung, C., van der Wall, B., Pausder, H., Burley, C., Brooks, T.,
Beaumier, P., Delrieux, Y., Mercker, E., and Pengel, K., “The HART-II Test:
Rotor Wakes and Aeroacoustics with Higher-Harmonic Pitch Control (HHC) Inputs -
The Joint German/French/Dutch/US Project,” in American Helicopter Society 58th
Annual Forum Proceedings, (Montreal, Canada), June 11-13, 2002.
[117] Zahle, F., Johansen, J., Sørensen, N. N., and Graham, J. M. R., “Wind
Turbine Rotor-Tower Interaction Using and Incompressible Overset Grid Method,”
in 45th AIAA Aerospace Sciences Meeting and Exhibit, (Reno, NV), AIAA-2007-0425,
Jan. 8-11, 2007.
[118] Zahle, F. and Sørensen, N. N., “Overset Flow Simulation on a Modern Wind
Turbine,” in 26th AIAA Applied Aerodynamics Conference, (Honolulu, HI), AIAA-
2008-6727, Aug. 18-21, 2008.
116
VITA
Charles Eric Lynch was born on Decemeber 5, 1981 in Atlanta, Georgia. He is married to
Eric developed an interest in flying machines at an early age. During a summer program
at Georgia Tech while he was still in elementary school, he obtained a course catalog and
was immediately fascinated reading the descriptions of courses with names like “High Speed
a Bachelor of Science in Aerospace Engineering from Georgia Tech with highest honors in
2004 and a Master of Science in Aerospace Engineering in 2007. It was during an internship
in the wind energy department at Sandia National Labs in Albuquerque, New Mexico in the
summer of 2006, that Eric first became interested in using CFD to simulate wind turbine
at Georgia Tech.
Eric is a member of the American Institute of Aeronautics and Astronautics and the
117