Vous êtes sur la page 1sur 12

Downloaded from SAE International by University of Michigan, Sunday, July 29, 2018

Lattice-Boltzmann Analysis of Three-Dimensional Ice Shapes 2015-01-2084

on a NACA 23012 Airfoil Published 06/15/2015

Benedikt König and Ehab Fares


Exa GmbH

Andy P. Broeren
NASA John H. Glenn Research Center

CITATION: König, B., Fares, E., and Broeren, A., "Lattice-Boltzmann Analysis of Three-Dimensional Ice Shapes on a NACA 23012
Airfoil," SAE Technical Paper 2015-01-2084, 2015, doi:10.4271/2015-01-2084.
Copyright © 2015 SAE International

Abstract The icing research community has known for many years that the
presence of ice accretion on any aerodynamic surface becomes a
A Lattice-Boltzmann approach is used to simulate the aerodynamics
challenging computational simulation problem almost immediately.
of complex three-dimensional ice shapes on a NACA 23012 airfoil.
Even in the initial stages of an exposure to icing conditions,
The digitally produced high fidelity geometrical ice shapes were
localized separated flow regions and boundary-layer transition may
created using a novel laser scanning technique in the NASA Icing
take place [2]. Continued exposure to these conditions results in
Research Tunnel. The geometrically fully resolved unsteady
larger ice accretion geometries and larger-scale separated flows. The
simulations are conducted on two ice shapes representing a roughness
glaze ice case with large leading-edge horn geometries is usually
type and a horn type icing on the leading edge of the airfoil.
considered to be the most challenging, owing to the large degree of
Comparisons between simulation and experiment of lift, drag, and
unsteady, three-dimensional separated flow. As described in Ref.
pitching moment as well as pressure distributions indicate overall a
[2], a similar situation exists for large spanwise-ridge type ice that
good qualitative agreement in capturing the aerodynamic degradation.
can form downstream of a wing leading edge. Numerous CFD
Especially for the horn-type ice shape, the quantitative agreement is
studies have been conducted with many different types of
also mostly very good. Analysis of the flow structures indicates
approaches including Reynolds-averaged Navier-Stokes (RANS),
furthermore a good capturing of the three-dimensional separation
Large Eddy Simulation (LES), Detached Eddy Simulation (DES)
behavior of the flow.
along with various combinations or “hybrid” schemes. The
geometric modeling of the ice geometry has also varied widely in
Introduction previous studies from simple two-dimensional representations to
fully three-dimensional geometries.
The use of computational tools for the design, development and
certification of aircraft continues to expand at a seemingly
Prior to the mid-2000's, most iced-airfoil CFD studies focused on
unyielding exponential rate. There is a widespread proliferation of
the implementation of 2D RANS methods for leading-edge and
CFD codes and related tools (such as for grid generation) that
spanwise-ridge ice. For example, Potapczuk [3] reasonably
provide engineers and researchers access to such tools in nearly
reproduced the experimental aerodynamics of an iced NACA0012
every part of the world. This situation is described in a recently
airfoil prior to the onset of fully separated flow associated with
released NASA report [1]. According to that report's authors, the
stall. Numerous parametric variations of simulated ice shape size,
advances in CFD capabilities realized over the last several decades
location, airfoil geometry, and Reynolds and Mach number have
has: led to significant reductions in both ground-based and in-flight
been carried out with 2D RANS simulations employing various
testing; reduced cost and program risk while providing superior
numerical schemes and turbulence models [4, 5, 6]. These results
designs; and provided deeper insight into the fundamental physics
typically have reasonable agreement with experimental data, or
of fluid dynamic behaviors heretofore unseen. In spite of these
predict the appropriate aerodynamic trends, in the range of lift
successes, the authors of Ref. [1] stress that current conventional
coefficient that is linear with angle of attack. Once flow separation
CFD methods still cannot reliably predict turbulent separated flows
becomes large, introducing the non-linearity, the level of agreement
and offer several recommendations to address this issue. The report
with experimental data degrades significantly. Thompson, et al. [7]
also provides numerous examples of such turbulent separated flows
extended a 2D RANS method to a 3D domain and included
for which improved methods are required.
spanwise geometry perturbations of the leading-edge ice shape.
However, the results were not significantly improved over 2D
results reported earlier.
Downloaded from SAE International by University of Michigan, Sunday, July 29, 2018

Because understanding iced-airfoil stall and the associated maximum were generated, and a subset of these was selected for aerodynamic
lift coefficient is an important flight-safety consideration, research evaluation. Table 1 provides the aerodynamic and icing cloud
over the past 10 years or so has focused on more sophisticated CFD conditions for the selected cases. The first column, “Ice-Shape
methods. Pan and Loth [8] implemented a 3D DES approach, but Classification” refers to the aerodynamic classification described by
were impeded in part by the larger computing resource requirements. Bragg et al [2]. For this investigation, one ice roughness case
Duclercq et al. [9] describe a hybrid RANS method called Zonal (ED1974) and the horn ice shape (ED1978) were selected since the
Detached Eddy Simulation (ZDES) that was applied to a large aerodynamic characteristics are significantly different between the
spanwise-ridge type ice accretion. The ice geometry was extruded in two cases. Thus each ice shape represented different flow simulation
the spanwise direction to complete the 3D CFD domain. The challenges for the CFD analysis. For the ice shapes listed in Table 1,
unsteady calculations were used to look at vortex shedding a 3D laser scan was performed to document the ice-accretion. The
frequencies for a single angle of attack case. It appears that the laser-scan data were used to fabricate artificial ice shapes using
computationally intensive nature of the simulation limited the number rapid-prototype manufacturing (RPM). Ref. [14] provides a detailed
of angles of attack that could be considered. Agreement of the mean description of the process used to develop the RPM artificial ice
results with experimental data were improved from standard RANS shapes from the laser-scan data. For both cases, the RPM shapes were
methods but results for more angles of attack are needed to fully manufactured using stereolithography. Data provided by the
assess the capability. Similar results were reported in Refs. [10] and manufacturer indicated an accuracy of ±0.015 inches with a minimum
[11] where the results of hybrid RANS/LES/DES methods were layer thickness of 0.005 inches. This procedure required generation of
compared against standard RANS results for various leading-edge a stereolithography file (*.stl) that contained the watertight ice-shape
glaze ice flow simulations. Both Refs. [10] and [12] were able to geometry. This same geometry file was used directly as input to the
create the 3D CFD meshes based upon measured 3D geometry of the CFD simulations. The spanwise length of the finished artificial ice
ice accretion. This capability is a relatively new development in icing shapes was limited to one-third of the aerodynamic model span, so
CFD simulations, however, the added geometric fidelity does not that three identical sections were required to cover the span of the
seem to have helped to improve the simulation results relative to model in the University of Illinois wind tunnel where subsequent
experimental data. More work is needed in this area to determine the aerodynamic performance tests were done.
level of geometric fidelity required to improve CFD simulations.
Graphical information for each of the selected cases is shown in
While the results of the latest unsteady, 3D hybrid RANS/LES Figure 1 and Figure 2. The section cuts were extracted from the
approaches for iced airfoil/wing configurations are encouraging, it is laser-scan data. For the RPM artificial ice shapes, the laser scan was
clear that further research is required to improve the overall acquired with the artificial ice shape bolted to the leading edge of the
confidence level in icing CFD. Given the challenge of simulating, NACA 23012 wind-tunnel model. Therefore, the laser-scan data
unsteady, separated and turbulent flow, Ref. [1] recommends a accurately represent the geometry as it was tested in the aerodynamic
number of alternative approaches to those identified in this literature wind tunnel. Each figure shows a 3D rendering of the artificial ice
review. One such novel nontraditional approach is the Lattice- shape solid model used for production of the RPM shapes. A
Boltzmann method which is based upon kinetic gas theory instead of photograph of the ice accretion generated in the IRT is also shown.
the Navier-Stokes equations. This method was previously used with
some success in comparison to RANS methods for a number of
iced-airfoil configurations [13]. The purpose of this paper is to
present and discuss the results of an improved Lattice-Boltzmann
method from Ref. [13]. Two 3D ice accretion geometries were
simulated on a NACA 23012 airfoil at conditions corresponding to
experimental aerodynamic data acquired at a Reynolds number of
1.8×106 and a Mach number of 0.18. The unsteady simulations were
performed on a fully three-dimensional domain based upon a
laser-scan geometry of the ice accretion and included modeling of the
wind-tunnel walls. Results comparisons were made for lift, drag,
pitching-moment and surface-pressure coefficients. Some of the CFD
flow field results were also compared to experimental flow
visualization images. A preliminary unsteady analysis compared two
dominant unsteady modes to previous experimental findings.

Ice-Accretion Geometry and Experimental


Aerodynamic Data
The ice-accretion geometry information and corresponding
experimental aerodynamic data utilized in this investigation were
originally acquired for the purposes of validating a commercial 3D
laser scanning system for recording ice-accretion geometry in the
NASA Icing Research Tunnel (IRT) [14, 15]. Ice-accretion tests were Figure 1. ED1974 roughness ice shape: section cut extracted from laser-scan
conducted in the IRT using an 18-inch chord, 2D straight wing with data (top), rendering of RPM artificial ice shape (bottom left), and ice-
accretion photograph (bottom right), after [15]
NACA 23012 airfoil section. Several different types of ice accretions
Downloaded from SAE International by University of Michigan, Sunday, July 29, 2018

surface. The RPM-based artificial ice shapes were also instrumented


with static pressure taps. The model lift and pitching-moment data
were acquired from a force balance and by integration of airfoil
surface static pressures measured by an electronically scanned
pressure system. Good agreement between these methods was
generally obtained. Locating and installing pressure orifices in and
around 3D ice features sometimes led to spurious pressure
coefficients on the artificial ice shape at certain angles of attack.
These spurious data could affect the lift and pitching-moment
coefficients determined from integrated surface pressure data. For
this reason, and for consistency with the results presented in [16], the
lift and pitching moment data for the clean configuration were
obtained from the surface pressures, whereas the data reported for the
iced configurations were obtained from the force balance in this
paper. Momentum-deficit methods were used to compute the drag
coefficient from total-pressure measurements collected by a
traversable wake rake. This way, contributions from the corner flow
at the model-wind tunnel junctions could be avoided. Surface-oil flow
visualization was also conducted for selected cases. Ref. [15]
provides more details about the experimental methodology, data
uncertainty, and wind-tunnel wall corrections.

Figure 2. ED1978 horn-ice shape: section cut extracted from laser-scan data Numerical Method
(top), rendering of RPM artificial ice shape (bottom left), and ice-accretion
The Lattice-Boltzmann Method, including its implementation in the
photograph (bottom right), after [15]
commercial software PowerFLOW as it was used for the simulations
The ED1974 roughness case in Figure 1 was a glaze-ice roughness presented in this paper, is based on kinetic gas theory. LBM is a CFD
with the typical “smooth zone” in the region of the stagnation point technology developed over the last 25-30 years [17, 18, 19, 20].
with large roughness features farther downstream. The ED1978
horn-ice case in Figure 2 was a typical glaze ice accretion with both
The Lattice-Boltzmann Approach
upper-and lower-surface horns. Refs. [14] and [15] provide graphical
In contrast to methods based on the Navier-Stokes (N-S) equations,
information of this type for the other ice accretion cases shown in
LBM is based on a simpler and more general physics formulation
Table 1.
[17]. Its motivation is to simulate a fluid at a microscopic level where
Table 1. Summary of IRT Test Conditions for Ice Shapes Selected for the physics are simpler and more general than the macroscopic
Aerodynamic Evaluation (after [15]). continuum approach taken by the N-S equations. However, as the
complete microscopic reproduction of molecular dynamics is
computationally much too expensive, a simplified mesoscopic
description is constructed.

Rather than tracking and solving for every molecule in a fluid, LBM
uses particle density distribution functions. In Kinetic theory, the
continuous distribution function describes the number of
particles at a given time t and position with a certain velocity .
From those distribution functions it is possible to obtain the
macroscopic quantities for density, momentum and energy by
All aerodynamic testing was performed using the low-speed, integration over the velocity space.
low-turbulence wind tunnel at the University of Illinois with the
experimental apparatus described by Broeren et al [15]. The wind • Density:
tunnel has a 33.6-in. (0.85-m) by 48-in. (1.2-m) test section capable
of speeds up to Mach 0.20. An 18-in. (0.46-m) chord NACA 23012
airfoil model was designed with interchangeable leading edges that
(1)
accommodated the various ice simulations. There was a baseline
leading edge having the NACA 23012 profile that was used to
• Momentum:
document the un-iced, or clean, airfoil performance, and there were
two ice leading edges with truncated nose geometry that allowed for
the attachment of the ice simulations. The artificial ice shapes bolted
onto the ice leading edges and thus had a rigid, repeatable mounting
system. The model had a main chordwise row of pressure taps, a (2)
secondary chordwise row, and a set of spanwise taps on the upper
Downloaded from SAE International by University of Michigan, Sunday, July 29, 2018

Now, the Boltzmann equation describes the rate of change of the Turbulence Modelling
velocity distribution function due to transport and collision LBM can generally be combined with various turbulence modeling
approaches ranging from a standard turbulence model, an LES
subgrid scale model, and also hybrid approaches. In the current
(3)
implementation, a modified k-ε two-equation model based on the
original RNG formulation describes the sub-grid turbulence
where the collision term C satisfies the conservation laws for mass, contributions [30] and is solved on the same lattice using a second
momentum and energy. By choosing the collision term appropriately, order scheme. A swirl correction for the local eddy viscosity is
it is possible to recover the macroscopic hydrodynamics of the fundamental in physically allowing the large vortical fluid structures
Navier-Stokes equations [21, 22]. From kinetic theory, the to develop and persist without artificial numerical damping. The swirl
discretization of the velocity space as well as of time and space leads model together with the inherently unsteady nature of the Lattice-
to the discrete Lattice-Boltzmann equation Boltzmann equation adequately reproduces the large scale turbulent
vortices. This represents, from a pragmatic point of view [31], a key
factor in predicting LES similar solutions on coarse grids using an
(4) unsteady turbulence model, a methodology referred to as Very Large
Eddy Simulation (VLES). As the model is explicitly independent of
The PowerFLOW solver employed in this paper thereby uses a the local cell size, it does not exhibit grid-induced separation
D3Q19 model which discretizes the continuous velocity space with behavior, as long as the base flow and the pressure gradients are
a set of 19 discrete velocities in three-dimensional space, illustrated captured adequately. Also, by not using an explicit sub-grid scale
in Figure 3. The collision operator C is simplified using the model, the VLES approach reduces resolution requirements
Bhatnagar-Gross-Krook (BGK) [22] approximation reproducing the compared to LES in regions where the flow can be captured by a
behavior of a wide range of fluids. It ensures the exact conservation RANS-like model. It thereby makes the industrial application of the
of local mass, momentum and energy. tool to complex high Reynolds number cases feasible [25].

This LBM-VLES based description of turbulent fluctuation carries


flow history and upstream information, and contains high order
terms to account for the nonlinearity of the Reynolds stress [32].
This is in contrast to Navier-Stokes based methods, which generally
use the conventional linear eddy viscosity based Reynolds stress
closure models.

Boundary Conditions and Wall Treatment


Inflow or outflow boundary conditions based on simple
Figure 3. Illustration of the three-dimensional D3Q19 model
extrapolations or simplified characteristics are easily defined using
the assumption of local equilibrium at the boundary. More
The dynamics of a fluid then consist of two steps, namely the complex non-reflecting unsteady boundary conditions are also
propagation from one lattice cell to another and the collision of the easily integrated.
particles within one cell. This leads to an inherently transient process
running at a CFL condition of exactly one and resembles an explicit The standard Lattice-Boltzmann bounce-back boundary condition for
time marching scheme. The solution process of evolving the velocity no-slip or the specular reflection for free-slip condition are
distribution functions is inherently parallel, as it only requires nearest generalized through a volumetric formulation [18] near the wall for
neighbor information, and is also stable. This makes for a arbitrarily oriented surface elements (Surfels) within the Cartesian
computationally efficient and robust method. The clear advantage that volume elements (Voxels). This formulation of the boundary
LBM methods offer is the very high temporal resolution, inherently condition on a curved surface cutting the Cartesian grid is
efficient unsteady simulation algorithm, and low dissipation and automatically mass, momentum and energy conservative while
dispersion of the numerical scheme [23]. For some cases [24, 25, 26, maintaining the general spatial second order accuracy of the
27], improvements in computational efficiency of one order of underlying LBM numerical scheme. This accurate formulation can be
magnitude were found compared to classical Navier-Stokes solutions. related to the kinetic foundation of the LBM theory, in contrast to the
immersed boundary approaches for the Navier-Stokes equations. For
The classical LBM scheme, as described here, is weakly the latter, the formulation of a conservative, accurate, and high-order
compressible and as such typically valid in the low Mach number wall boundary condition remains a challenging task, which makes it
regime up to around Mach 0.4. Recent extensions to the scheme [28, difficult to exploit the benefits Cartesian meshes can offer in an
29], however, allow the simulation of flows at higher Mach numbers accurate manner.
as well.
In order to reduce the resolution requirements near the wall for high
Reynolds number flows, a hybrid wall function is used to model the
wall-nearest part of the boundary layer on solid surfaces. The wall
function model [33] is an extension of the standard log-law of the
wall, including the laminar sublayer and the buffer layer, as well as
Downloaded from SAE International by University of Michigan, Sunday, July 29, 2018

the effect of favorable and adverse pressure gradients. It accounts for boundary condition to avoid the development of a wall boundary
surface roughness through a length parameter. Various validations of layer. It is described above that each ice shape model only covered
the wall model, especially the accurate prediction of the separation one third of the model span in experiment and, thus, was repeated
point on a smooth body such as stall of a wing have been published three times. To limit the computational effort of the simulations, the
previously [34, 35]. setup was restricted to one third of the actual span as well. The flow
conditions were set according to the aerodynamic performance tests
done at the University of Illinois to Mach 0.18 and a Reynolds
Computational Grids number of 1.8 million.
With the Cartesian mesh approach it is possible to develop automatic
meshing tools that can handle arbitrarily complex geometries. Local In the experiments, non-dimensionalization was done based on
grid refinement is achieved by splitting voxels with a factor of two reference quantities measured 48 inches upstream in the wind tunnel,
uniformly in each direction. Different to many other Cartesian grid whereas the CFD simulations used the nominal quantities. To account
methods, the current implementation uses the aforementioned surfel for the small deviation between the two, all CFD pressure
concept rather than an immersed boundary technique. This allows for distributions presented here are corrected for the ΔCp between the
an accurate and robust representation of complex surfaces. The nominal condition and the one measured at the reference position.
method has, for example, been repeatedly successfully applied to This correction was found to be 0 ≤ ΔCp < 0.1 for all simulations. Lift
simulations of complete aircraft configurations including high-lift and pitching moment measurements in CFD were based on surface
devices and deployed landing gears [25]. integration on the complete model, similar to the balance
measurements in the wind tunnel. For drag, the corner flow effects
Grid resolution for aerodynamic simulations is generally driven by were avoided by limiting the integration area on the model to only
three aspects. In the near-wall region, a resolution of five to ten 5% of the model span in the center. This approach is somewhat
percent of the boundary layer thickness (corresponding to 100 ≲ y+ ≲ comparable to using the wake survey method in the experiments. The
300) is typically used but the three-layer wall model can also handle a comparisons to measurements are based on un-corrected wind tunnel
much wider range. Close to the simulated body, the resolution is such data as the inclusion of the wind tunnel test section in the simulations
as to resolve the smallest geometrical structures of interest. Similarly, should account for most of the interference effects.
the resolution in the flow volume should be fine enough to resolve the
relevant coherent flow structures. Two different types of ice shapes were simulated using the classical,
weakly compressible Lattice-Boltzmann method described above. Its
An illustration of the Cartesian mesh around the three-dimensional validity for the current application was verified a posteriori by
horn ice shape investigated in this work is given in Figure 4. The confirming that local flow velocities in the simulations stayed within
volume resolution here is of the same size as the smallest geometrical the acceptable range of the method. The two ice shapes were
structures represented in the ice shape. Roughness 1 (ED1974, see Figure 1) and the Horn shape (ED1978,
see Figure 2), according to Table 1. The stereolithography files
described above were directly used as input into the numerical model,
i.e. no simplifications were applied and no labor-intensive
rediscretization was needed. The finest resolution of the modelled
surface hence corresponds to the resolution of the stereolithography
files provided from the laser scan data. The finest cell size in the flow
volume was chosen such as to resolve the smallest structures of the
horn ice shape ED1978 with at least one voxel. This implies that the
smallest flow structures are under-resolved in the simulations. The
larger and more relevant structures are resolved sufficiently fine, as
will be shown by the grid resolution study below. For the roughness
ice shape ED1974, the smallest voxel size was set to resolve a
roughness element of average height by at least four voxels. The
Figure 4. Illustration of the volume grid around the ED1978 horn ice shape overall grid sizes were N = 33.8 × 106 and N = 23.0 × 106 voxels for
(grid coarsened for illustration purposes)
ED1978 and ED1974, respectively.

The Numerical Setup A grid convergence study was conducted to assess the impact of the
The numerical setup included the baseline airfoil NACA 23012 resolution on the results for the ED1978 horn ice shape. Three setups
installed in a wind tunnel test section with dimensions based on the were run, coarse, medium and fine, with successive refinement ratios
University of Illinois wind tunnel described above. Wind tunnel of 1.25 each. The behavior of the lift coefficient with resolution is
walls were simulated to account for potential blockage effects of the presented in Figure 5, where a grid resolution index N−2/3 is used as a
flow separations of varying size for the different ice shapes and flow measure for the average grid spacing squared, according to [36]. For
conditions. To simplify the setup, a constant cross section was second order accurate schemes, the result should form a straight line.
chosen rather than the slightly diverging walls of the actual wind The resolution study shows that the medium resolution,
tunnel. Diverging walls are used in the wind tunnel to mitigate corresponding to the one used in this work, is well within the grid
streamwise pressure gradients along the test section due to the convergent range. By extrapolating the line in Figure 5 towards an
growing boundary layer on the walls. For the purpose of the infinitesimally fine grid with N−2/3 → 0 it can be seen that the current
simulations conducted herein, the walls were modeled with a slip medium resolution overestimates lift by about ΔCL ∼0.02 due to the
Downloaded from SAE International by University of Michigan, Sunday, July 29, 2018

finite resolution. It should be noted, however, that exact quadratic


grid convergence (i.e. a straight line in Figure 5) cannot necessarily
be expected to be achieved for flows with large scale flow separation,
as is the case for this horn ice shape.

Figure 5. Grid convergence behavior of lift versus a grid resolution index [36]

Results
The two ice shapes, ED1974 and ED1978, were selected for the
initial code validations presented in this work as they cover the
two extremes in terms of ice shapes and their impact on
aerodynamic performance.

Figure 6. Aerodynamic performance comparison for the ED1974 roughness


NACA 23012 - Clean Airfoil ice shape
Results of the clean airfoil are shown in the following sections
together with the ice shape results. From Figure 6 it can be seen that The pitching moment in Figure 6 follows a similar trend as the lift
the clean simulation matches the experimental lift curve well within coefficient. Again, the range of lower angles of attack is well captured
the linear range, but it under-predicts the non-linearity in the lift slope by the numerical simulation for both the clean and the iced airfoils.
close to stall. Related to that is the same under-prediction of the At higher angles of attack, the non-linear change in the polar is again
nonlinearity of the pitching moment polar at high angles of attack. It not captured. In the experimental results, both configurations develop
is assumed that the leading edge stall, as a consequence of a laminar a slight pitch-up tendency as they approach stall. This suggests that
separation bubble on the NACA 23012, is not reproduced correctly in the lift reduction in the non-linear region of the lift curve is caused by
the simulation. This could be related to the resolution near the leading a reduction in lift in the rear part of the airfoil. It will be shown in the
edge or to the wall model applied in conjunction with the fully discussion of the pressure distribution later that this is related to a
turbulent boundary layer assumption. This also explains the higher small trailing edge separation.
drag in the simulations of the clean airfoil compared to the
experiments that featured some laminar flow. The drag polar, shown at the bottom of Figure 6, confirms that the ice
roughness effects are under-predicted. While the divergence of drag
is qualitatively well reproduced, there is a substantial under-
Ice Roughness prediction of the absolute levels.
The first ice shape investigated is the roughness ice shape denoted
ED1974. An illustration of this ice shape was given in Figure 1. Besides the global characteristics of the force polars it is also
Forces and pitching moment results for this ice shape are compared interesting to consider the incremental changes caused by the ice
to the clean airfoil in Figure 6. As was mentioned before, there are roughness. This can help to cancel out the discrepancies seen already
some discrepancies already in the polars of the clean airfoil, where in the clean airfoil simulations. Figure 7 shows the difference in lift
the non-linear behavior at high angles of attack is not reproduced between the iced and the clean airfoil, for both the wind tunnel and
correctly by the numerical simulation. The same trend is visible for the CFD simulations. From this incremental plot it is clear to see that
the simulation of the iced airfoil. The non-linear effect of the ice on the effect of the ice shape on the lift performance is well captured,
the lift curve prior to stall is captured qualitatively but the reduction with only a slight shift of about one degree to higher incidences.
in lift slope is slightly under-predicted. The overall polar shape,
however, is reproduced well. In particular, the angle of attack for For a more detailed understanding of how experiment and numerical
maximum lift matches the experimental value within one degree. The simulation compare, a pressure distribution at the maximum lift
stall characteristic is similar for both simulation and experiment, with condition for the iced airfoil is presented in Figure 8. Considering the
a very gradual, trailing edge-type stall. The over-prediction of clean airfoil first, the pressure distribution shows a very good
maximum lift is largely due to the combination of a higher predicted agreement between the wind tunnel measurement and CFD. Only on
lift-slope in the non-linear range and the slightly delayed stall. the lower surface is the pressure slightly too high in the numerical
simulation, which is in line with the slight over-prediction of lift. For
the iced airfoil, the agreement is qualitatively still very good. The
reduction in suction pressure due to the ice roughness from the
Downloaded from SAE International by University of Michigan, Sunday, July 29, 2018

leading edge up to x/c ≈ 0.2 is reproduced, albeit not as pronounced


as in the wind tunnel. Again, this is in agreement with the relative
over-prediction of lift in CFD.

Figure 7. Lift increment between iced and clean condition for ED1974

Figure 9. Aerodynamic performance comparison for the ED1978 horn ice


shape

The pitching moment also shows very good agreement to the


Figure 8. Pressure distribution comparison for the ED1974 roughness ice experimental results in the region just up to stall. Around the
shape at α=10° maximum lift condition, the pitch characteristic becomes more
nose-down compared to the experiment. The qualitative agreement is,
In Broeren et al [15] it is highlighted that the divergence of Cp however, still good. Drag, shown at the bottom of Figure 9, is again
downstream of x/c ≈ 0.85 between the clean airfoil and the iced one in very good agreement with the experiments, both qualitatively and
in the wind tunnel is related to a trailing edge separation. While this quantitatively. Looking at the lift increment due to the ice shape,
divergence in Cp is visible in the CFD results as well, no trailing edge shown in Figure 10, confirms that the effects of the ED1978 horn ice
separation is predicted. This lack of a separation could be the reason shape on the airfoil's lift performance are very well captured by the
for the stronger nose-down pitching moment seen in CFD. In a numerical simulation.
previous study with the same CFD method [35] it was shown that a
trailing edge separation in quasi two-dimensional wind tunnel
experiments may be triggered by an interaction with the side wall
boundary layer. With the tunnel side walls modeled with slip
boundary conditions in this study there is no such trigger and this
may explain why the CFD solution is not able to capture the trailing
edge separation.

Horn Ice
The second ice shape investigated numerically in this work is the Figure 10. Lift increment between iced and clean condition for ED1979 horn
ED1978 horn ice. An illustration of this ice shape was given in Figure ice shape
2. It represents the opposite extreme from the roughness ice presented The comparison of the pressure distributions for α = 6°, presented in
before with massive flow separations in the wakes of the ice horns. Figure 11, also highlights the good overall agreement. The distinctive
Again, forces and moment polars are shown in Figure 9 for the clean plateau region on the upper surface, just aft of the leading edge, is
airfoil and for the ice shape. The comments made before concerning accurately reproduced. A similar feature on the lower surface leading
the clean airfoil apply here as well. The lift curve of the iced airfoil edge is also matching the experimental results well. The lower
shows a reasonably good match to the experimental values. Both, the surface overall is showing again a slightly too high pressure, similar
lift slope and the early degradation in lift are well captured. The angle to what was reported previously. On the upper surface, downstream
of attack for maximum lift is also well predicted, again within one of the plateau, the initial pressure recovery shows some minor
degree of the wind tunnel result. Maximum lift CL,max itself is differences to the experiments. The plateau itself and this first part of
over-predicted by ΔCL ∼ 0.1. In the post-stall region, the agreement the pressure rise, up to approximately x/c ≈ 0.25, are related to the
is again very good. The reasons for the deviation in CL,max will be recirculation area behind the ice horn. The steeper pressure rise in
discussed later on. that area indicates a slightly smaller recirculation region in the CFD
simulations. This will also be confirmed later in the flow
Downloaded from SAE International by University of Michigan, Sunday, July 29, 2018

visualization. This difference in the pressure gradient is then quickly denote the approximate locations of the separation bubble
reduced and the trailing edge pressure levels are again very well reattachment. The qualitative agreement between the two
matched between CFD and wind tunnel. Overall, the agreement at reattachment lines and their spanwise variation is very good,
this pre-stall angle of attack is very good. indicating that the CFD simulation was able to properly capture all
relevant flow structures from the complex ice shape. Quantitatively,
the reattachment location was slightly too far upstream in the
numerical simulation. This is consistent with the conclusion drawn
from the pressure distribution in Figure 11.

Figure 11. Pressure distribution comparison for the ED1978 horn ice shape at
α = 6°

To get an overview of the pressure distributions at different flow


conditions, Cp for a number of angles of attack is plotted in Figure
12. This confirms the very good agreement of the numerical
simulations at lower angles of attack. In the stalling region, here
shown for α = 8° and 10°, there are certain discrepancies which seem
Figure 13. Surface flow visualization from experiment (left side, oil flow,
to be related to the shape and extent of the upper surface flow
adapted from [37] with permission) and time-averaged CFD (right side, skin
separation. The differences to the experimental distributions seem to
friction lines) for α = 6°. Yellow lines denote reattachement lines.
indicate an effectively increased thickness and camber of the airfoil.
This could be a result of the CFD simulation predicting a wider and The data presented so far were all based on time-averaged results. To
longer recirculation region on the upper surface. It is clear, however, fully appreciate the nature of the flow, it is also interesting to consider
that the pressure distributions at α = 8° and 10° are consistent with instantaneous conditions and compare them to the time-averaged
the differences between the experimental and numerical lift and ones. Figure 14 shows a comparison of instantaneous (top) and
moment polars discussed before. In the post-stall regime, here shown time-averaged (bottom) distributions of the velocity magnitude in the
at α = 12°, the airfoil is massively separated and the numerical center plane for the ED1978 horn ice shape at an angle of attack of α
simulation agrees again well with the experiments. Overall, this is = 6°. The instantaneous snapshot shows a wealth of smaller and
showing some differences of the numerical method at stall condition, larger structures that are not captured in the time-averaged solution. It
but an otherwise very good agreement to the wind tunnel data. particularly highlights the break-up process of the shear layer
emanating from the upper horn. These results are consistent with the
experimental results found in [38, 39] for a NACA0012 airfoil with
leading-edge horn-ice shape. Furthermore, the basic unsteady versus
time-averaged view of the large-scale separated flow post horn-ice is
comparable to the literature review of Ref. [2].

Figure 12. Overview of pressure distributions for the ED1978 horn ice shape
at various angles of attack.

For an even more detailed flow analysis, Figure 13 compares the


surface flow structures between the experiments and the numerical
simulations at α = 6°. On the left hand side an oil-flow visualization,
adapted from Monastero [37], is shown. The right hand side shows
skin friction lines from the CFD simulation. The original oil-flow Figure 14. Illustration of instantaneous (top) and time-averaged (bottom, with
picture was trimmed to show the middle one of the three identical ice streamlines) distributions of velocity magnitude for the ED1978 horn ice
shape sections that were installed in the wind tunnel. The yellow lines shape for α = 6°.
Downloaded from SAE International by University of Michigan, Sunday, July 29, 2018

Also interesting to note is the high degree of unsteadiness in the flow and the accompanying discussion, identifying the mean reattachment
that is already present at this pre-stall angle of attack. The time location is not a precise exercise and therefore contributes some
histories of the lift coefficient in Figure 15 further show how the uncertainty in the resulting Strouhal numbers. That said, the resulting
unsteadiness of the flow quickly increases when the stall condition at values were St ≈ 0.3 and 0.6 for α = 6° and α = 8°, respectively. The
around α = 8° is reached. The time-averaged streamlines in the lower latter value of St ≈ 0.6 is well within the range identified by Gurbacki
part of Figure 14 illustrate the separation bubble in the wake of the and Bragg [39] for their iced-airfoil case as well as values cited in the
upper horn in shape and size. Its extent agrees well with the surface literature for other separated flow geometries such as a backward-
visualizations shown in Figure 13. facing step. While further investigation of the unsteady flow field for
these CFD simulations is required, this preliminary analysis has
shown the potential for the CFD simulations to accurately reproduce
the unsteady characteristics observed in past experiments.

Besides the unsteadiness mentioned above, another important feature


of the flow around such an ice shape is the complexity and three-
dimensionality of its topology. Figure 17 depicts iso-surfaces of total
pressure, again for an instantaneous snapshot (top) and a time-
averaged solution (bottom). The iso-surfaces are color-coded with the
velocity magnitude shown previously in Figure 14. The instantaneous
Figure 15. Time history of lift coefficient for the CFD simulations of the snapshot at the top gives an impression of the complexity of both the
ED1978 horn ice shape at selected angles of attack. underlying ice shape and the resulting flow. The streamwise position
of the shear layer breakup varies significantly over the span of the
The positive qualitative comparisons of the unsteady features in the three-dimensional horn ice shape. This information gets smoothed out
iced-airfoil flow field near stall with the experimental results of Refs. when only time-averaged results (bottom) are considered. To be able
[38] and [39] warranted some preliminary analysis of the unsteady to capture this complexity it is important to apply a numerical method
content. For the numerical simulations at the pre-stall angle of attack that offers a high accuracy both in space and time together with low
of α = 6°, and at the maximum lift condition at α = 8°, a longer time dissipation properties, in order to properly capture the fine structures
signal was collected to perform a spectral analysis of the lift visible here.
coefficient, shown in Figure 16. For both angles of attack, there are
two peaks in the spectra, one at a low frequency around 10 Hz and
another at a higher frequency near 130 Hz.

Figure 16. Power spectral density of the lift coefficient.

Gurbacki and Bragg [39] also identified similar frequency content in


their experimental study of a horn ice shape/NACA0012 airfoil flow
field. The low-frequency oscillation was characterized as a larger-
scale separation and reattachment of the upper surface flow and has
been related to a phenomenon called “shear-layer flapping.” Gurbacki
and Bragg calculated a Strouhal number using the airfoil projected
height (c sin α) as the length scale and the freestream velocity as the
velocity scale. Applying this definition of Strouhal number to the
frequency peaks near 10 Hz in Figure 16 yields St ≈ 0.01 which
compares very favorably to the data compiled in Ref. [39].
Figure 17. Iso-surfaces of total pressure (Cp,tot = −1.3) of instantaneous (top)
The higher frequency peak near 130 Hz was characterized as a and time-averaged (bottom) flow for ED1978, colored with velocity
“regular vortex shedding” mode of the separated shear layer. In this magnitude.
case, Gurbacki and Bragg calculated a Strouhal number using the
mean reattachment length of the separation bubble as the length scale. To assess the quality of the results presented here it is useful to draw
The freestream speed was again used as the velocity scale. For the a comparison to current state-of-the-art numerical simulations. Some
present simulations, the mean reattachment length was approximated limited results on the same horn ice shape were presented by Jun et al
at x/c = 0.3 for α = 6° and x/c = 0.6 for α = 8°. As shown in Figure 13 [12]. The focus of that work was on a feasibility study to investigate
Downloaded from SAE International by University of Michigan, Sunday, July 29, 2018

if existing grid generation and RANS-CFD tools have the potential to flow were shown around stall. These limitations were also observed
simulate those complex ice shapes. Jun et al could successfully create for the absolute performance results of the subtle ice-roughness
surface-fitted grids and conduct their unsteady RANS simulations. shape, where the differences to the clean airfoil aerodynamics were
The amount of data presented, however, is not sufficient for a relatively small. The incremental effect was well captured, though.
meaningful comparison with the current results. For the horn ice shape, which had a less subtle effect on the airfoil
flow and performance, the simulations agreed very well with the
Instead, some numerical studies on different glaze-ice shapes [8, 10] experimental results, both incrementally and in absolute terms. For
will be used for comparison. In the most recent study, Brown et al both ice shapes considered, the stalling angle of attack is captured by
[10] presented Implicit Large Eddy Simulations (ILES) of a the numerical simulation to within one degree of the experiments.
comparable, high-fidelity three-dimensional ice shape geometry. The lift degradation due to the ice shapes is also well predicted for
Besides reporting considerable difficulties in creating a computational the two cases, despite their significantly different aerodynamic
mesh, they conclude that the flow dynamics captured by their ILES characteristics. Some limitations related to the stalling mechanism of
approach improved the results compared to RANS simulations. The the clean airfoil exist and need to be further investigated and
comparison to experimental lift data still seems less favorable for the understood. A preliminary unsteady analysis showed good agreement
ILES than the data presented in this work. It is also interesting to note of two distinct modes, one low- and one high-frequency, to previous
that Brown et al report the runtimes of their simulations between one experimental results.
to two days on 960 cores, corresponding to 23-46 kCPUh per data
point. This compares to 4-12 kCPUh for the simulations at α = 4° and The overall quality of the results suggests that the Lattice-
6° shown in Figure 15, respectively. Boltzmann method, together with the Very Large Eddy Simulation
technology employed here, is capable of sufficiently predicting the
In an earlier study, Alam et al [11] conducted RANS and hybrid aerodynamic performance of complex three-dimensional ice shapes.
RANS/LES simulations on an extruded two-dimensional horn-ice It was shown that the method can handle the highly detailed
shape, albeit only for one single angle of attack. While the reported geometries obtained from the three-dimensional laser scans and it is
lift coefficient agrees reasonably well with the experimental value for expected that similar simulations of such ice shapes can be
most of the tools investigated, there are significant deviations in the performed on full aircraft configurations.
pressure distribution shown. No angle of attack is reported in that
study, but the shape of the pressure distribution seems to suggest that
a flow condition prior to stall was simulated. All four N-S methods in References
[11] failed to predict the suction level on the upper surface and the 1. Slotnick J., Khodadoust A., Alonso J., Darmofal D., Gropp W.,
size of the recirculation region correctly. The simulations at pre-stall Lurie E. and Mavriplis D., “CFD Vision 2030 Study: A Path to
conditions presented here seem to correlate very well with the Revolutionary Computational Aerosciences,” NASA Langley
pressure measurements in Figure 11. Research Center, 2014.
2. Bragg M., Broeren A. and Blumenthal L., “Iced-Airfoil
The comparison of the LBM method presented herein to some recent Aerodynamics,” Progress in Aerospace Sciences, 41(5):323-
N-S simulations is, due to the lack of sufficiently comparable data, 418, July 2005, doi:10.1016/j.paerosci.2005.07.001.
more incidental than of a rigorous nature. The comparisons do,
3. Potapczuk M. G., “Numerical analysis of an NACA 0012
however, provide some indication that the LBM technology is a
airfoil with leading-edge ice accretions,” Journal of Aircraft,
promising candidate to improve the capability for numerical
25(3):193-194, 1988, doi:10.2514/3.45576.
performance simulations of ice shapes on airfoils and wings.
4. Dunn T., Loth E. and Bragg M., “Computational Investigation
of Simulated Large-Droplet Ice Shapes on Airfoil
Summary Aerodynamics,” Journal of Aircraft, 36(5):836-843, 1999,
High-fidelity numerical simulations of laser-scanned three- doi:10.2514/2.2517.
dimensional ice shapes using a Lattice-Boltzmann approach were 5. Kumar S. and Loth E., “Aerodynamic Simulations of Airfoils
presented. A number of different types of ice shapes on the NACA with Upper Surface Ice Shapes,” Journal of Aircraft, 38(2):285-
23012 airfoil had previously been measured and scanned at the 295, 2001, doi:10.2514/2.2761.
NASA Icing Research Tunnel. Two ice shapes with different 6. Pan J. and Loth E., “Reynolds-Averaged Navier-Stokes
aerodynamic characteristics were selected for the current study. The Simulations of Airfoils and Wings with Ice Shapes,” Journal of
numerical method employed allowed to directly incorporate the very Aircraft, 41(4):879-891, 2004, doi:10.2514/1.587.
detailed stereolithography data into the numerical model, without the
7. Thompson D., Mogili P., Chalasani S., Addy H. and Choo Y.,
need for simplifications or labor intensive re-discretization of the ice
“A Computational Icing Effects Study for a Three-Dimensional
shape geometries.
Wing,” in 42nd AIAA Aerospace Sciences Meeting and Exhibit,
AIAA Paper 2004-561, 2004, doi:10.2514/6.2004-561.
The Lattice-Boltzmann numerical method used in the commercial
software PowerFLOW was shown to capture the incremental effects 8. Pan J. and Loth E., “Detached Eddy Simulations for Iced
of the ice shapes very well, based on comparisons of integral forces Airfoils,” Journal of Aircraft, 42(6):1452-1461, 2005,
and local pressure distributions. For absolute performance doi:10.2514/1.11860.
predictions, the quality of the results ranged between very good and
reasonable, depending on the underlying flow physics. Some
limitations of the current simulations in predicting the clean airfoil
Downloaded from SAE International by University of Michigan, Sunday, July 29, 2018

9. Duclercq M., Brunet V. and Moens F., “Physical Analysis of 22. Qian Y. H., D'Humières D. and Lallemand P., “Lattice BGK
the Separated Flow around and Iced Airfoil based on ZDES Models for Navier-Stokes Equation,” Europhysics Letters,
Simulations,” in 4th AIAA Atmospheric and Space Environments 17:479-484, 1992, doi:10.1209/0295-5075/17/6/001.
Conference, AIAA Paper 2012-2798, 2012, doi:10.2514/6.2012- 23. Shan X., Yuan X.-F. and Chen H., “Kinetic theory representation
2798. of hydrodynamics:a way beyond the Navier-Stokes equation,”
10. Brown C., Kunz R., Kinzel M., Lindau J., Palacios J. and Journal of Fluid Mechanics, 550: 413-441, 2006, doi:10.1017/
Brenter K., “RANS and LES Simulation of Airfoil Ice S0022112005008153.
Accretion Aerodynamics,” in 6th AIAA Atmospheric and Space 24. Zhang R., Shan X. and Chen H., “Efficient kinetic method for
Environments Conference, AIAA Paper 2014-2203, 2014, fluid simulation beyond the Navier-Stokes equation,” Physical
doi:10.2514/6.2014-2203. Review E, 74: 046703, 2006, doi:0.1103/PhysRevE.74.046703.
11. Alam M., Thompson D. and Walters D., “Hybrid Reynolds- 25. Marié S., Ricot D. and Sagaut P., “Comparison between lattice
Averaged Navier-Stokes/Large-Eddy Simulation Methods for Boltzmann method and Navier-Stokes high order schemes
Flow Around and Iced Wing,” Journal of Aircraft, 52(1):244- for computational aeroacoustics,” Journal of Computational
256, 2015, doi:10.2514/1.C032678 Physics, 228:1056-1070, 2009, doi:10.1016/j.jcp.2008.10.021.
12. Jun G., Oliden D., Potapczuk M. and Tsao J.-C., “Computational 26. Manoha E., “Category 5 Results Summary: ONERA/Airbus
Aerodynamic Analysis of Three-Dimensional Ice Shapes on LAGOON Simplified Landing Gear configuration,” in Third
a NACA 23012 Airfoil,” in 6th AIAA Atmospheric and Space AIAA Workshop on Benchmark Problems for Airframe Noise
Environments Conference, AIAA Paper 2014-2202, 2014, Computations, 2014.
doi:10.2514/6.2014-2202.
27. Khorrami M., Fares E. and Casalino D., “Towards Full Aircraft
13. Chi X., Li Y., Chen H., Addy H., Choo Y. and Shih T.-P., Airframe Noise Prediction: Lattice Boltzmann Simulations,” in
“A Comparative Study Using CFD to Predict Iced Airfoil 20th AIAA/CEAS Aeroacoustics Conference, AIAA Paper 2014-
Aerodynamics,” in 43rd Aerospace Sciences Meeting and 2481, 2014, doi:10.2514/6.2014-2481.
Exhibit, AIAA Paper 2005-1371, 2005, doi:10.2514/6.2005-
28. Khorrami M. R. and Mineck R. E., “Towards Full Aircraft
1371.
Airframe Noise Prediction: Detached Eddy Simulations,” in
14. Lee S., Broeren A., Kreeger R., Potapczuk M. and Utt 20th AIAA/CEAS Aeroacoustics Conference, AIAA paper 2014-
L., “Implementation and Validation of 3-D Ice Accretion 2480, 2014, doi:10.2514/6.2014-2480.
Measurement Methodology,” in 6th AIAA Atmospheric and
29. Lockard D. P., “Summary of the Tandem Cylinder Solutions
Space Environments Conference, AIAA Paper 2014-2613, 2014,
from the Benchmark problems for Airframe Noise
doi:10.2514/6.2014-2613.
Computations-I Workshop,” in 49th AIAA Aerospace Sciences
15. Broeren A., Addy H. J., Lee S. and Monastero M., “Validation Meeting, AIAA paper 2011-353, 2011, doi:10.2514/6.2011-353.
of 3-D Ice Accretion Measurement Methodology for
30. Yakhot V. and Orszag S., “Renormalization Group Analysis of
Experimental Aerodynamic Simulation,” in 6th AIAA
Turbulence,” Journal of Scientific Computing, 1(2):3-51, 1986,
Atmospheric and Space Environments Conference, AIAA Paper
doi:10.1007/BF01061452.
2014-2614, 2014, doi:10.2514/6.2014-2614.
31. Menter F., Kuntz M. and Bender R., “A Scale Adaptive
16. Broeren A., Addy H. J., Bragg M., Busch G., Guffond D. and
Simulation Model for Turbulent Flow Predictions,” in 41st
Montreuil E., “Aerodynamic Simulation of Ice Accretion on
Aerospace Sciences Meeting and Exhibit, AIAA Paper 2003-
Airfoils,” NASA/TP-2001-216929, 2011.
0767, 2003, doi:10.2514/6.2003-767.
17. Chen H., “Volumetric Formulation of the Lattice-Boltzmann
32. Chen H., Orszag S., Staroselsky I. and Succi S., “Expanded
Method for Fluid Dynamics: Basic Concept,” Physical Review
Analogy between Boltzmann Kinetic Theory of Fluid and
E, 58(3):3955-3963, 1998, doi:10.1103/PhysRevE.58.3955.
Turbulence,” Journal of FLuid Mechanics, 519:307-314, 2004,
18. Chen H., Texeira C. and Molvig K., “Realization of FLuid doi:10.1017/S0022112004001211.
Boundary Condition via Discrete Boltzmann Dynamics,”
33. Fares E., “Unsteady Flow Simulation of the Ahmed Reference
Int. Journal of Modern Physics C, 9(8):1281-1292, 1998,
Body using a Lattice Boltzmann Approach,” Journal of
doi:10.1142/S0129183198001151.
Computers and Fluids, 35(8):940-950, 2006, doi:10.1016/j.
19. Chen H., Kandasamy S., Orszag S., Shock R., Succi S. and compfluid.2005.04.011.
Yakhot V., “Extended Boltzmann Kinetic Equation for Turbulent
34. Noelting S., Fares E. and Keating A., “Simulations of the
Flows,” Science, 301(5633):633-636, 2003, doi:10.1126/
Trapwing Case with PowerFLOW,” in HiLiftPW-1 Workshop,
science.1085048.
2010.
20. Chen S. and Doolen G. D., “Lattice Boltzmann Method for
35. König B., Fares E., Noelting S., Jammalamadaka A. and Li Y.,
Fluid Flows,” Annual Review of Fluid Mechanics, 30:29-364,
“Investigation of the NACA 4412 Trailing Edge Separation
1998, doi:10.1146/annurev.fluid.30.1.329.
using a Lattice-Boltzmann Approach,” in 44th AIAA Fluid
21. Chen H., Chen S. and Matthaeus W. H., “Recovery of the Dynamics Conference, AIAA Paper 2014-3324, 2014.
Navier-Stokes equations using a lattice-gas Boltzmann method,”
36. Mavriplis D. J., Vassberg J. C., Tinoco E. N., Mani M.,
Physical Review A, 45(8):R5339-R5342, 1992, doi:10.1103/
Brodersen O. P., Eisfeld B., Wahls R. A., Morrison J. H.,
PhysRevA.45.R5339.
Zickuhr T., Levy D. and Murayama M., “Grid Quality and
Resolution Issues from the Drag Prediction Workshop Series,”
Journal of Aircraft, 46(3):935-950, 2009, doi:10.2514/1.39201.
Downloaded from SAE International by University of Michigan, Sunday, July 29, 2018

37. Monastero M. C., “Validation of 3-D Ice Accretion 42. König B., Fares E. and Noelting S., “Lattice-Boltzmann Flow
Measurement Methodology Using Pressure-Sensitive Paint,” Simulations for the HiLiftPW-2,” in 52nd Aerospace Sciences
Master thesis, University of Illinois at Urbana-Champaign, Meeting, AIAA paper 2014-0911, 2014, doi:10.2514/6.2014-
2013. 0911.
38. Jacobs J. and Bragg M., “Two- and Three-Dimensional Iced 43. König B., Fares E. and Noelting S., “Fully-Resolved Lattice-
Airfoil Separation Bubble Measurements by Particle Image Boltzmann Simulation of Vane-Type Vortex Generators,” in 7th
Velocimetry,” in 45th AIAA Aerospace Sciences Meeting and AIAA Flow Control Conference, AIAA paper 2014-2795, 2014,
Exhibit, AIAA paper 2007-88, 2007, doi:10.2514/6.2007-88. doi:10.2514/6.2014-2795.
39. Gurbacki H. and Bragg M., “Unsteady Flowfield About an Iced 44. Vatsa V. N., Casalino D., Lin J. C. and Appelbaum J.,
Airfoil,” in 42nd AIAA Aerospace Sciences Meeting and Exhibit, “Numerical Simulation of a High-Lift Configuration with
AIAA paper 2004-562, 2004, doi:10.2514/6.2004-562. Embedded Fluidic Actuators,” in 32nd AIAA Applied
40. Larsson J. and Wang Q., “The prospect of using LES and DES Aerodynamics Conference, AIAA paper 2014-2142, 2014,
in engineering design, and the research required to get there,” doi:10.2514/6.2014-2142.
Royal Society Philosophical Transactions A, 372(2022), 2014,
doi:10.1098/rsta.2013.0329.
41. Larsson J. and Kawai S., “Wall-modeling in large eddy
simulation: length scales, grid resolution and accuracy,” in
Annual Research Briefs, Center for Turbulence Research, 2010.

The Engineering Meetings Board has approved this paper for publication. It has successfully completed SAE’s peer review process under the supervision of the session organizer. The process
requires a minimum of three (3) reviews by industry experts.

All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or transmitted, in any form or by any means, electronic, mechanical, photocopying, recording, or
otherwise, without the prior written permission of SAE International.

Positions and opinions advanced in this paper are those of the author(s) and not necessarily those of SAE International. The author is solely responsible for the content of the paper.

ISSN 0148-7191

http://papers.sae.org/2015-01-2084

Vous aimerez peut-être aussi