You are on page 1of 14

AIAA Atmospheric and Space Environments Conference AIAA 2010-7985

2 - 5 August 2010, Toronto, Ontario Canada

Prediction of Rotor Blade Ice Shedding using Empirical Methods

Jeremy Bain, Juan Cajigas, Lakshmi Sankar


School of Aerospace Engineering
Georgia Institute of Technology, Atlanta, GA 30332-0150

Robert J. Flemming
Sikorsky Aircraft Co., Stratford, CT

Roger Aubert
Downloaded by GEORGIA INST OF TECHNOLOGY on June 27, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.2010-7985

Bell Helicopter Textron Inc., Fort Worth, TX

ABSTRACT
A methodology that couples computational fluid dynamics, computational structural
dynamics, ice accretion models and ice shedding models is developed and applied to both
isolated airfoils and rotors in forward flight, the latter with and without shedding. The individual
modules are coupled to each other through industry-standard open file I/O methods, allowing the
replacement of the individual modules with more advanced modules as technology matures. Ice
shape results are presented and correlated with test data for a range of icing conditions. The
torque rise associated with ice build-up on a UH-60A rotor in forward flight is modeled. Finally,
results are presented for ice shedding phenomena for a small-scale model rotor. Reasonable
correlation with test data is observed in the cases studied.

BACKGROUND
Despite decades of research on the phenomenon, rotor icing remains a major in-flight
concern for many civilian and military helicopter operators. One particular facet of rotor blade
icing receiving additional attention is shedding, more specifically self shedding. Self shedding
occurs on a rotor blade when the aerodynamic and centrifugal forces on a section of ice exceed
the structural adhesion forces holding the ice onto the blade. At the point when the adhesion
force is exceeded, the ice is said to “shed” and separates from the rotor blade. After separation,
the shed ice acts as a projectile and has the potential to strike components of the helicopter such
as the tail rotor.
In order to better evaluate the risks associated with self shedding, an accurate model must be
developed for determining the conditions under which ice shedding will occur. This paper
focuses on self-shedding mechanics and the development of a model, based on a combination of
computational and empirical methods, for predicting shedding phenomena. The model presented
is being developed as part of a larger initiative to investigate the trajectories of shed ice pieces
and their potential for striking vehicle components. In order to meet this larger objective, the
present model will first be used to predict shedding characteristics at particular operating
conditions. At each operating condition analyzed, CFD and six degree-of-freedom modeling
would then be used to determine the trajectories of shed ice pieces, as well as possible impact
forces.

Copyright © 2010 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.
NUMERICAL FORMULATION
Ice accretion on helicopter rotors is a process that depends on many physical variables – blade
geometry, number of blades, operating conditions (thrust setting, hub moments, advance ratio,
rotor RPM), ambient temperature, air pressure, liquid water content, and the time period over
which accretion occurs. In the present study, these conditions were specified at the start of the
simulation. The specific values and conditions chosen are discussed later in the paper.
The present approach for prediction of ice shedding phenomena is a multi-disciplinary
methodology that involves aerodynamics, structural dynamics, rotor trim, ice accretion, and
shedding. The overall methodology is shown in Figure 1. The individual pieces that form this
collection of tools are discussed below.
Downloaded by GEORGIA INST OF TECHNOLOGY on June 27, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.2010-7985

Clean rotor
Specify CT/σ, advance
ratio, desired hubloads PYTHON Framework
LEWICE3D
CFD / CSD Transform flow-field to
Run LEWICE3D at specified
Loose Coupling blade fixed coordinate
azimuthal intervals, add new
Trimmed solution system
layer of ice to existing shapes

CGT
Convert iced airfoil data
Build computational grid
into rotor geometry files
for iced rotor

Trajectory Analysis Shedding Analysis

Figure 1. Overview of the Ice Accretion, Shedding and Trajectory Analysis

LEWICE3D: LEWICE3D is a grid-based ice accretion software analysis tool that can interface
with a variety of 3D computational solvers for computing ice shapes on three-dimensional
external surfaces. The streamlines and ice particle trajectories are computed using the provided
flow solution. The ice accretion analysis is done over the blade surface that is divided into 2D
sections-of-interest. A 2D heat transfer module is used to calculate the ice growth along the
streamline. LEWICE3D and previous versions have been validated for a wide range of airfoils
and icing conditions.1,2 although mostly for fixed wing aircraft as opposed to rotating wing/blade
associated with rotorcraft.

CFD: The Navier-Stokes computational fluid dynamics code OVERFLOW Version 2.1y was
used for the 3D rotor simulations. OVERFLOW, developed by NASA, uses overset structured
grids for a wide variety of problems including rotorcraft aerodynamic simulations.3,4

2
In some simulations, a hybrid Navier-Stokes/free wake solver called GT-Hybrid was used.
GT-Hybrid is a three-dimensional unsteady viscous compressible flow solver. The flow is
modeled from first-principles using the Navier-Stokes methodology. The three-dimensional
unsteady Navier-Stokes equations are solved in the transformed body-fitted coordinate system
using a time-accurate, finite volume scheme. A third-order spatially accurate Roe scheme is used
for computing the inviscid fluxes and second order central differencing scheme for viscous
terms. The Navier-Stokes equations are integrated in time by means of an approximate LU-
implicit time marching scheme. SA-DES turbulence model is used to compute the eddy
viscosity. The flow is assumed to be turbulent everywhere, and hence no transition model is
currently used.
A single blade is resolved through the Navier-Stokes solution for faster and economical
computations. The influence of the other blades and of the trailing vorticity in the far field wake
Downloaded by GEORGIA INST OF TECHNOLOGY on June 27, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.2010-7985

is accounted for by modeling them as a collection of piece-wise linear bound and trailing vortex
elements. The near wake is captured inherently in the Navier-Stokes analysis. The use of such a
hybrid Navier-Stokes/ vortex modeling method allows for an accurate and economical modeling
of viscous features near the blades, and an accurate “non-diffusive” modeling of the trailing
wake in the far field.5,6
Because of the modular nature of the present coupling methodologies, either of these solvers
(OVERFLOW or GT-Hybrid) may be used based on user’s experience and preference. The CFD
solver communicates with the icing analysis using standard PLOT3D format flow field and grid
information. It communicates with the structural analysis using standard Fluid-Structure-
Interaction files.

Computational Structural Dynamics (CSD): In the present framework, any industry standard
comprehensive analysis may be used that exchanges with the CFD solver the grid motion, grid
deformation, and airloads in a predefined format. In the present study, DYMORE7 is used. This
is a finite element based solver that is used for structural dynamics analysis. The solver can
handle multi-body dynamics, and may be used for the analysis of complex geometrical
configurations with arbitrary topologies. For example, each component in a rotor system
including the blade, hinges, hub and pitch link may be modeled as separate elements, with their
connectivity modeled as constraint equations. The rotor blades are modeled as elastic beams with
a geometrically-exact composite beam finite element formulation. DYMORE is also considered
a rotorcraft comprehensive solver because it provides an internal aerodynamic model and
autopilot trimmer which may be used to do a trimmed aeroelastic rotorcraft analysis. The
aerodynamic model is a lifting line-based analysis with table lookup of aerodynamic coefficients.
A dynamic inflow model is used to compute the downwash.

CFD-CSD Loose Coupling Methodology: Rotorcraft aeromechanical studies involve coupling


the rotor aerodynamics with the structural dynamics of the system. The airloads computed by the
CFD solver is used to drive a forced response simulation with the CSD solver. The computed
structural deflections are used in the CFD analysis, leading to a change in the airloads. The two
solvers are thus inherently coupled. The CFD-CSD coupling may be performed primarily in two
ways – loose and tight. In tight coupling, the data is exchanged every time step of the simulation.
In loose coupling, the data is exchanged between the two solvers at periodic intervals, typically
once per revolution. Since loose coupling is driven by the inherent periodicity in the solution, it
is used for analysis of rotors in steady flight conditions.

3
The coupling methodology framework is shown in Figure 1. The first step involves running
the CSD code that computes airloads using its internal lifting line-based aerodynamic model.
These airloads are applied on the rotor structural model to compute the elastic deformations. The
solver also trims the rotor to match the measured hub loads by adjusting the pitch controls. The
periodic blade deformations obtained from this run are transferred to the CFD solver using a
fluid structure interface. The CFD solver deforms the blade mesh and computes the periodic
airloads, which are subsequently transferred to the structural dynamics system. The coupling
iterations are executed until convergence is observed in hub loads obtained from the CFD solver
and pitch controls obtained from the CSD solver.

Empirical Model for Shedding An empirical model for self-shedding was created by expanding
Downloaded by GEORGIA INST OF TECHNOLOGY on June 27, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.2010-7985

on the basic methods developed by Fortin et al8 and Flemming et al.9The following procedure
used:

1. At any specified instant in time, the contact area, volume, and mass of the ice are
computed. This is done using the simultaneous integration of the flow equations,
structural dynamics equations, and the ice accretion equations in time.
2. The shear stress at the blade surface between the ice mass and the blade and the cohesive
stresses exerted on a segment of ice by the neighboring ice mass are computed. The
surface shear stresses are based on temperature and on the rotor blade surface ,using
relationships derived from experimental data.
3. The components of the centrifugal, shear, and cohesive force vectors are summed up, on
sections on the rotor blade.
4. The feasibility of shedding is examined. It is assumed that all the ice mass outboard of a
given radial location will be shed if the sum of applied forces (centrifugal, edge
cohesion, and optionally aerodynamic pressure) on the mass of ice exceeds the adhesion
force.

VALIDATION OF THE COUPLED METHODOLOGY FOR ISOLATED AIRFOILS

Prior to the application of the present methodology to the shedding of ice from rotor
configurations, the coupling methodology shown in Figure 1 was tested. Four separate validation
cases for a NACA 0012 airfoil were considered, with the experimental data being supplied from
Flemming et al9 and Shin et al.10 Table I below shows the flow and icing conditions for each of
the validation cases.

Table I. Icing conditions for NACA 0012 validation


AoA LWC Time Diameter
Run Mach T (K)
(deg) (g/mm3) (s)
411 0.29 0 263.15 0.96 60 20
2077 0.29 6 263.15 0.3 60 20
404 0.32 4 256.49 0.55 420 20
Shin 0.21 4 247.04 1.0 360 20

Each of these cases is at relatively low Mach number, and experiment has shown that these
conditions produce relatively smooth ice shapes. The methodology outlined above was tested

4
with the loop shown in Figure 1, executed a variable number of times. Because only a 2D non-
rotating airfoil configuration was being studied, the structural dynamics solver, the trim process,
the ice shedding model, and the trajectory module were not invoked. The calculation loop may
be executed just once with LEWICE3D being run a single time in a one-shot approach for the
total icing simulation time (“Couple 1x”). When the calculation loop is executed more than once
(“Couple 2x,” Couple 3x,” etc.), LEWICE3D is run several times, each for a fraction of the total
simulation time, with an updated flow field provided for each iteration. Figure 2 shows the
LEWICE3D predictions for ice shape for the four validation cases shown in Table I.

0.08 0.08

0.06 0.06
Downloaded by GEORGIA INST OF TECHNOLOGY on June 27, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.2010-7985

0.04 0.04

0.02 Clean 0.02 Clean


Experiment Experiment
0 0
Couple 1x Couple 1x
-0.02 -0.02
Couple 3x Couple 3x
-0.04 -0.04

-0.06 -0.06

-0.08 -0.08
-0.05 0 0.05 0.1 0.15 -0.05 -0.03 -0.01 0.01 0.03 0.05 0.07 0.09 0.11 0.13 0.15

Nondimensional Chord Length Nondimensional Chord Length

Figure 2a. NACA 0012 validation. Run 411 on the left, Run 2077 on the right

0.08 0.08

0.06 0.06

0.04 0.04
Clean
0.02 0.02 Clean
Experiment
Experiment
0 Couple 1x 0
Couple 1x
-0.02
Couple 3x -0.02 Couple 2x
Couple 6x
-0.04 -0.04

-0.06 -0.06

-0.08 -0.08
-0.05 -0.03 -0.01 0.01 0.03 0.05 0.07 0.09 0.11 0.13 0.15 -0.1 -0.05 0 0.05 0.1 0.15

Nondimensional Chord Length Nondimensional Chord Length

Figure 2b. NACA 0012 validation. Run 404 on the left, Shin (Ref. 8) on the right.

In each run, the multiple time step coupling provides a more accurate ice shape in terms of
impingement location, total ice volume, or both. From the results shown in Figure 2, it can be
seen that the methodology described above provides only qualitatively reasonable results.
Additional work is in progress to improve the correlation between measured and computed data
for ice shapes, which may subsequently be used in shedding analyses.

5
VALIDATION OF HOVERING ROTOR COMPUTATION

The high fidelity aerodynamic simulations were done using the compressible Navier-Stokes
Computational Fluid Dynamics code OVERFLOW version 2.1y.11 OVERFLOW, originally
developed by NASA, uses overset structured grids for a wide variety of problems including
rotorcraft simulations.12,13 Since hover and axial flight do not have an azimuthal dependency, the
flow around one blade was computed and periodic boundary conditions were applied to account
for the other blades. This allows the problem to be computed in steady state with a single blade
greatly reducing the total computational effort. To evaluate the boundary conditions and grid
requirements, a UH-60A grid was created using Chimera Grid Tools.14 The main blade grid is an
O-grid of dimension 101x73x85 in the chordwise, spanwise, and normal directions. The surface
grids are shown in Figure 3. A center hub similar to one used by Strawn and Djomehri15 was
Downloaded by GEORGIA INST OF TECHNOLOGY on June 27, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.2010-7985

used to prevent the root vortices from going above the rotor plane. The first point normal to the
wall was placed at y+=1 and the blade grid extended one chord in the normal direction. The off
body grids are cylindrical with spacing of 0.02 tip chords in tip region. The domain extends 5
radii above, 20 below and 10 out from the center of rotation and is shown in Figure 4. The total
grid consisted of 8.7 million points and computational time was 36 hours on 8 processors. The
use of a single blade allows much greater grid resolution for the same computational time as
standard forward flight helicopter simulations. Axial flight also reduces the control variables to
just collective and eliminates the need to trim the rotor to desired hub moments.

Figure 3. UH-60A OVERFLOW Surface Grids with Every Second Point in Both Directions
Shown

6
Downloaded by GEORGIA INST OF TECHNOLOGY on June 27, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.2010-7985

Figure 4. UH-60A Off Body Grid Layout

For the UH-60A rotor in hover, there have been several scaled rotor tests, full-scale wind
tunnel tests, and flight tests. These tests are summarized by Shinoda et al16 and show a
significant variation in Figure of Merit on the order of 0.05, especially at higher thrusts. Two
scale model tests using a 1 to 5.73 scale model were tested by Lorber in 1989 and 1991 in two
different wind tunnels.17,18 A full-scale hover test in the NASA Ames 80- by 120-foot NFAC
wind tunnel, using production UH-60A blades, was made at low to moderate thrust values.16
Flight tests have also been made but show a very wide range in Figure of Merit due to
uncertainty in ambient conditions and fuselage download.19 Numerous computational efforts
using these data sets have been done to improve computational performance and wake trajectory
calculations. These include four blade coupled CFD/CSD12, single blade full Navier-Stokes15,20,21
and single blade hybrid Navier-Stokes and vorticity embedding potential flow.22 The
performance calculations for a range of collective values are shown in Figure 5. The computed
data shows excellent correlation with the NFAC data for low to medium thrust values. (The
experiment was not conducted at high thrust values due to wind tunnel wall effects.) The
computed results show a consistently lower Figure of Merit than the scaled wind tunnel tests. No
attempts were made to correct for scale model effects and including these would result in
improved performance for the model rotors, further increasing the discrepancy. The rotor was
assumed to be rigid with a two-degree precone and any elastic twist was neglected. The elastic
deformation of one to two degrees for the full-scale blade at high thrust condition would result in
increased twist and improved performance. The model rotors were heavily instrumented
resulting in increased blade weight, lower coning, and different dynamic properties.

7
0.80

0.70

0.60 OVERFLOW
Lorber 89

FM
Lorber 91
80x120
0.50

0.40

0.30
0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10 0.11 0.12
CT/σ
σ
Downloaded by GEORGIA INST OF TECHNOLOGY on June 27, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.2010-7985

Figure 5. Computed UH-60A Hover Results Compared to Several Experiments

VALIDATION OF THE COUPLED METHODOLOGY FOR


ROTORS IN FORWARD FLIGHT

The methodology shown in Figure 1 has also been applied to the 3D analysis of a rotor in
forward flight. However the 3D problem is much more complicated than 2D isolated airfoil
analysis, due to the flexibility of the rotor. Aeroelastic effects must be taken into account, so a
computational structural dynamics solver must be used as well. As stated earlier, the CSD code
DYMORE was be coupled to a CFD analysis, GT-Hybrid. This was done in a loosely coupled
manner where the CFD and CSD solvers exchange information once after every revolution.
In the simulations presented here a two-step ice shape update process was used, meaning that
LEWICE3D was invoked twice, at both 150 sec and at 300 sec after the CFD code reached
steady solution. First the CFD/CSD coupling was employed on the clean rotor to compute the
airloads and the flow field. The flow field information at 0°, 90°, 180°, and at 270° azimuth was
fed into LEWICE3D. At each of these azimuths LEWICE3D was run for a simulation time of
one quarter of the total icing time step (150 sec). LEWICE3D analysis was done over the rotor
separately analyzing 2D slices of the rotor and marching from the blade root to the blade tip.
When LEWICE3D updated the ice shape, the CFD/CSD coupling was resumed and the iced
rotor was retrimmed to meet target hub loads. This concludes step one of the two-step process –
the first 150 seconds of flight. This process was repeated for a second time, and the simulations
were stopped at a total ice accretion time of 300 seconds.
The present simulations are for the steady level flight of a UH-60A, flight 76-2, documented
in NACA CR 391023. The total increase in engine torque was measured. The advance ratio for
this forward flight condition is 0.21. Table II shows the icing flight conditions for this analysis.

Table II. Icing flight conditions for forward flight analysis.

Gross TAS Altitude Static LWC Dd Time Torque


Weight (kg)(m/s) (m) Temp ( C)(g/m3) (s) Rise
780 46 690 -11 0.25 30 300 6%

8
As stated earlier, the total icing time is 300 seconds, and each step of the two-step process
simulated 150 seconds of ice accretion. The overall goal is to accurately predict the torque rise.
An increase in rotor torque due to icing requires the engine to supply more power to maintain
steady level flight. An assessment of the torque rise allows rotorcraft designers and analysts to
determine the capability of the rotorcraft in icing conditions.
Since the x-force contributes to the torque, it makes sense to look at the azimuthal variation in
the x-force as the ice accretion process advances in time. The results at the end of 150 sec and
300 sec are shown in Figure 6.

1.5E-04

1.0E-04
Downloaded by GEORGIA INST OF TECHNOLOGY on June 27, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.2010-7985

Clean
5.0E-05
Step 1
0.0E+00 Step 2
-5.0E-05

-1.0E-04

-1.5E-04

-2.0E-04

-2.5E-04
0 50 100 150 200 250 300 350
Azimuth
Figure 6. Azimuthal variation in x-force for flight 76-2.

The rise in torque as a result of ice accretion was also computed after 150 sec and 300 sec. Table
III shows the predicted torque rise after each step of the ice accretion update process.
Interestingly, the torque rise after step two (at the end of 300 sec) is less than after step one (after
150 sec). This is not as surprising when the x-force data from Figure 6 is analyzed. There are
many azimuthal locations where the x-force after step one is greater than the x-force after step
two. It was observed that a more aerodynamic shape forms over the rotor blade as the ice builds
from step one to step two. This may lead to a decrease the torque rise. The uneven variation in
ice formation and torque variation also affirms the fact that ice accretion is a complex nonlinear
process.

Table III. Predicted torque rise with IACM two-step method.


% Increase
in Torque
MR Only: Step 1 5.4%
MR Only: Step 2 4.9%
Aircraft Observed 6.0%

9
It is noted is that the present method under predicts the aircraft observed engine torque rise by
about 18% (4.9% compared to 6%), but this can be explained by the fact that the present method
only considers the rise in main rotor torque, whereas the experimental data is measuring engine
torque for the complete helicopter. The main rotor torque rise is responsible for the majority of
the rotorcraft’s contribution to the engine torque rise, but not all of it. If this analysis were to be
incremented for the effect of icing on other subsystems, such as the fuselage or tail rotor, it is
expected that the present method would come closer to the 6% engine torque rise.

APPLICATION OF THE PRESENT METHOLOGY TO


ICE SHEDDING PHENOMENA

Following the application of the present coupling methodology to isolated airfoils and rotors
Downloaded by GEORGIA INST OF TECHNOLOGY on June 27, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.2010-7985

in forward flight, calculations were done for modeling the ice shedding phenomena. A rotor
configuration tested by Fortin and Ferron (Ref. 8) was considered. The rotor is a 1/18-scale
model of a small helicopter. The rotor diameter is 0.78 m, and the tip speed is 130 m/s. The
forward speed of the rotor was 15 m/sec, leading to a low advance ratio (forward speed to tip
speed ratio) of 0.115. The blades are untwisted, and made of NACA 0012 sections (0.07 m in
chord), with a root cut out of 0.075 m. The liquid water content was 0.84 g/m3, and the median
diameter of the water drops was 27 µm. The ambient temperature was parametrically varied
between -20 deg Celsius and -5 degree C.
Since the model rotor was rigid, it was not necessary to employ computational structural
dynamics simulations. Furthermore, the rotor was operated at a fixed collective pitch of 6
degrees with zero cyclic. Therefore, it was not necessary to re-trim the rotor as the ice build-up
and gradually altered the aerodynamic loads.
The power required to operate the rotor was monitored both in the experiments and
simulations. As the ice began to build up, the rotor power gradually rose from 120 watts to 5200
watts. An abrupt drop in the power indicted that a shedding event had occurred. The time at
which shedding occurred and the approximate length and mass of the shed ice were recorded
using a video camera.
The simulations in the present work were done using OVERFLOW coupled to LEWICE3D.
After each update of the ice shape, a shedding analysis was done to determine if and when the
centrifugal forces outboard of a given radial station exceed the surface adhesion forces that exist
at each cross section of the ice shape. The accretion time at which such shedding occurs as well
as the thickness and length of the shed ice shape was extracted from the present simulations.
Figures 7 through 10 show comparisons with test data. The present simulations overestimated the
ice thickness over the entire length of the blade for several ambient temperatures. Reasonably
good agreement was found for the other properties, such as the length of the shed ice and the
time at which shedding occurs.

10
Downloaded by GEORGIA INST OF TECHNOLOGY on June 27, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.2010-7985

Figure 7. Ice Thickness along Blade


180

Experiment
160
Predicted

140
Shed Ice Length (mm)

120

100

80

60

40

20

0
-25 -20 -15 -10 -5 0
Temperature (C)

Figure 8. Comparison of Computed and Measured Lengths of the Shed Ice for
Test Conditions Considered in Ref. 8

11
250

Experiment
Predicted
200

Shedding Time (s)


150

100

50
Downloaded by GEORGIA INST OF TECHNOLOGY on June 27, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.2010-7985

0
-25 -20 -15 -10 -5 0
Temperature (C)

Figure 9. Comparison of Computed and Measured Time at which Shedding Occurs for
Test Conditions Considered in Ref. 8
16

14 Experiment
Predicted

12
Ice Thickness (mm)

10

0
-25 -20 -15 -10 -5 0
Temperature (C)

Figure 10. Effect of Ambient Temperature on Computed and Measured ice Thickness

CONCLUDING REMARKS

A coupled CFD + structural dynamics + trim + ice accretion + ice shedding + 6 DOF
trajectory analysis has been developed. The individual components are modular, and are coupled
to each other through file I/O. This allows the flow solver to be switched (from OVERFLOW to
GT-Hybrid for example), and the CSD solver to be switched (from DYMORE to a different
solver such as RCAS or CAMRAD-II). This modularity allows solvers of different texture to be
coupled and employed for the important problem of ice accretion on rotors and its effect on
power, torque, and shedding. Preliminary results have been presented for isolated airfoils, rotors
in forward flight, and for ice shedding phenomena to demonstrate that the coupling process is

12
robust and can be applied to a wide class of problems. Additional validation of the methodology
is needed to establish the reliability and usefulness of this methodology.

ACKNOWLEDGEMENTS
The coupling methodology shown in Figure 1 was developed under NASA NRA Cooperative
Agreement NNX08AU71A. Richard Kreeger was the technical monitor. The shedding modeling
and analysis was funded by the Center for Rotorcraft Innovation (CRI) and the National
Rotorcraft Technology Center (NRTC), U.S. Army Aviation and Missile Research, Development
and Engineering Center (AMRDEC) under Technology Investment Agreement W911W6-06-2-
0002, entitled National Rotorcraft Technology Center Research Program. The authors would like
to acknowledge that this research and development was accomplished with the support and
Downloaded by GEORGIA INST OF TECHNOLOGY on June 27, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.2010-7985

guidance of the NRTC and CRI. The views and conclusions contained in this document are those
of the authors and should not be interpreted as representing the official policies, either expressed
or implied, of the AMRDEC or the U.S. Government. The U.S. Government is authorized to
reproduce and distribute reprints for Government purposes notwithstanding any copyright
notation thereon. (Note that CRI is now known as the Vertical Lift Consortium or VLC.)

REFERENCES
1
Wright, W. B., “Validation Results for LEWICE 3.0," NASA/CR-2005-213561, March 2005.
2
Wright, W., “Further refinement of the LEWICE SLD Model,” AIAA Paper 2006-0464.
3
Buning, P.G., Chiu, I.T., Obayashi, S., Rizk, Y.M. and Steger, J.L., “Numerical Simulation of
the Integrated Space Shuttle Vehicle in Ascent,” AIAA Paper 88-4359.
4
Chan, W. M. “The OVERGRID Interface for Computational Simulations on Overset Grids,”
32nd AIAA Fluid Dynamics Conference, St. Louis, Missouri, June 2002. AIAA 2002-3188.
5
Nischint Rajmohan, Lakshmi Sankar, Olivier Bauchau, Bruce Charles, Stephen M. Makinen
and T. Alan Egolf, "Application of Hybrid Methodology to Rotors in Steady and Maneuvering
Flight,” AHS 64th Annual Forum, 2008.
6
Rajmohan, N., Manivannan, V., Sankar, L. N., Costello, M., and Bauchau, O., “Development
of a Methodology for Coupling Rotorcraft Aeromechanics and Vehicle Dynamics to Study
Helicopters in Maneuvering Flight,” AHS 65th Annual Forum, May 2009.
7
Bauchau O.A. and Kang, N.K.: “A Multi-Body Formulation for Helicopter Structural Dynamic
Analysis,” Journal of the American Helicopter Society, 38, No 2, April 1993.
8
Fortin, G and Perron, J., “Spinning Rotor Blade Tests in Icing Wind Tunnel,” AIAA Paper
2009-4260.
9
Flemming, R.J. and Lednicer, D.A., “High Speed Ice Accretion on Rotorcraft Airfoils,” NASA
CR-3910, November 1984.
10
Shin, J. and Bond, T. H., “Results of an icing test on a NACA 0012 Airfoil in the NASA
Lewis Icing Research Tunnel,” AIAA Paper 1992-0647.
11
Nichols, R., Tramel, R., and Buning, P., “Solver and Turbulence Model Upgrades to
OVERFLOW 2 for Unsteady and High-Speed Applications,” 24th AIAA Applied Aerodynamics
Conference, San Francisco, CA June 5-8, 2006, AIAA-2006-2824.
12
Potsdam, M., Yeo, H., Johnson, W., “Rotor Airloads Prediction using Loose
Aerodynamic/Structural Coupling,” American Helicopter Society 50th Annual Forum, Baltimore,
Maryland, June 7-10, 2004.
13
13
Boyd, D., “HART-II Acoustic Prediction using a Coupled CFD/CSD Method,” American
Helicopter Society 54th Annual Forum, Grapevine, Texas, May 27-29, 2009.
14
Chan, W., “The OVERGRID Interface for Computational Simulations on Overset Grids.”
32nd AIAA Fluid Dynamics Conference and Exhibit, St. Louis, Missouri, June 24-26, 2002,
AIAA-2002-3188.
15
Strawn, R. C., and Djomehri, M. J., “Computational Modeling of Hovering Rotor and Wake
Aerodynamics,” Journal of Aircraft, Vol. 39, (5), October 2002, pp. 786–793.
16
Shinoda, P., Yeo, H., and Norman, T., “Rotor Performance of a UH-60 Rotor System in the
NASA Ames 80- by 120-Foot Wind Tunnel,” American Helicopter Society 58th Annual Forum
Montreal, Canada June 11-13, 2002.
17
Lorber, P.F., Stauter, R.C., and Landgrebe, A.J., “A Comprehensive Hover Test of the
Downloaded by GEORGIA INST OF TECHNOLOGY on June 27, 2014 | http://arc.aiaa.org | DOI: 10.2514/6.2010-7985

Airloads and Airflow of an Extensively Instrumented Model Helicopter Rotor,” American


Helicopter Society 45th Annual Forum, Boston, MA, May 1989.
18
Lorber, P.F., “Aerodynamic Results of a Pressure- Instrumented Model Rotor Test at the
DNW,” Journal of the American Helicopter Society, Vol. 36, (4), October 1991.
19
Bousman, W.G., “Out-of-Ground-Effect Hover Performance of the UH-60A,” UH-60 Airloads
Program Occasional Note 2001-01, February 2001, http://rotorcraft.arc.nasa.gov/research/pdfs/
2001-01.pdf.
20
Wake, B. E., and Baeder, J. D., “Evaluation of a Navier– Stokes Analysis Method for Hover
Performance Prediction,” Journal of the American Helicopter Society, Vol. 41, (1), January
1996, pp. 7–17.
21
Vasilescu, R., Yeshala, N., Sankar, L. N., and Egolf, T. A., “Structured Adaptive Mesh
Refinement (SAMR) Algorithms Applied to Rotor Wake Capturing,” American Helicopter
Society 63rd Annual Forum Proceedings, Virginia Beach, VA, May 1–3, 2007.
22
Schmitz, S., Bhagwat, M., Moulton, M., Caradonna, F., and Chattot, J., “The Prediction and
Validation of Hover Performance and Detailed Blade Loads,” Journal of the American
Helicopter Society 54, 032004 (2009).
23
Flemming, R. J., and Lednicer, D. A., “High Speed Ice Accretion on Rotorcraft Airfoils,”
NASA CR 3910, 1985.

14

You might also like