You are on page 1of 14

AIAA 2008-432

46th AIAA Aerospace Sciences Meeting and Exhibit


7 - 10 January 2008, Reno, Nevada

Results From a Full-Scale Propeller Icing Test


Christopher Dumont 1
FAA William J. Hughes Technical Center, Atlantic City International Airport, NJ, 08405, USA
Paul Pellicano 2
FAA Atlanta Certification Office, Atlanta, Georgia, 30349, USA
Timothy Smith and James Riley 3
FAA William J. Hughes Technical Center, Atlantic City International Airport, NJ, 08405, USA

Tests to investigate and document leading-edge and runback propeller ice accretions in
Appendix C icing conditions, as well as an attempt to approximately duplicate a propeller ice
accretion from a Mitsubishi MU-2B flight test propeller which experienced full-span ice
accretion in suspected supercooled large drop (SLD) icing conditions were conducted in the
McKinley climactic chamber, Eglin AFB, in November 2006. This paper reviews the test
objectives, setup, and resulting data.

Nomenclature
AoA
hp
KTAS
LWC
MVD
OAT
RPM
SLD
VDC
Ac
0
n0
K0

=
=
=
=
=
=
=
=
=
=
=
=
=

angle of attack,
horse power
knots true air speed
liquid water content, g/m3
median volumetric diameter, m
outside air temperature, F
revolutions per minute
supercooled large drop
Volt direct current
accumulation parameter
collection efficiency
freezing fraction
modified droplet inertia parameter

I. Introduction
This paper provides an overview of a propeller icing wind tunnel test conducted to investigate and document ice
accretions on a full-scale propeller. The test was conducted in the McKinley Climatic Laboratory, at Eglin Air Force
Base, Florida. It was a collaborative effort of the Federal Aviation Administration (FAA) Aircraft Icing Research
program, FAA Aircraft Certification Service, United States Air Force (USAF), Hartzell Propeller, Inc., MT
Propeller Company, Goodrich Corporation, McCauley Propeller, and Hamilton Sundstrand. Several other people
and companies were instrumental in completing this work. They include Aerospace Testing Engineering &
Certification (AeroTEC), Intercontinental Jet Services Incorporated L.L.C., RS Information Systems (RSIS), Dr.
Andy Broeren, and Dr. David Anderson.

Aircraft Icing Research Program, AJP-6350


Icing Specialist for Small Airplane Directorate, ACE-111
3
Aircraft Icing Research Program, AJP-6350
2

1
American Institute of Aeronautics and Astronautics
This material is declared a work of the U.S. Government and is not subject to copyright protection in the United States.

During a recent flight test in an MU-2B in natural icing conditions, the aircraft experienced unexpected fullspan propeller icing. During the icing encounter, the airspeed decayed dramatically, losing 40 knots in 1 minute and
25 seconds1. Airframe ice accretion was deemed to be small by the flight crew. A postflight analysis of in-flight
video showed ice accretion along the entire span of the propeller blades. See Fig. 1. In-service experience of one
transport airplane indicated that propeller runback ice was a significant problem for a certain outside air temperature
(OAT) range, requiring redesign of the propeller ice protection system2.
Very little information on propeller icing is available in the public sector. The primary sources of experimental
data are still the papers by Preston and Blackman published in 19483 and Neel and Bright published in 19504. Both
of these reports document propeller performance losses for a research aircraft flying in natural icing conditions. Ice
accretion characteristics are not documented. No data from controlled testing in a ground icing facility is available.

Figure 1. In-flight video from MU-2B test flight showing


ice accretion along entire span of propeller blade.
The FAA partnered with industry to study this problem, and as a result, a test was done in an attempt to achieve
the following objectives:
1.

2.
3.
4.
5.

Document propeller ice accretions in Appendix C icing conditions.


Leading-edge accretions - size and shape, span location, and shedding frequency as a function of
icing conditions, blade material, blade condition, and revolutions per minute (RPM).
Runback ice accretions - size, shape, and location as a function of icing conditions, blade material,
blade condition, RPM, and deicing schedule and power.
Document propeller ice accretions in supercooled large drop (SLD) icing conditions and determine if a
propeller ice accretion from a Mitsubishi MU-2B flight test propeller in suspected SLD icing conditions
can be approximately duplicated.
Determine differences in thrust between noniced and iced propellers.
Evaluate differences between reference and scaled conditions.
Determine if nominal propeller efficiencies reductions can be proposed for certification.

2
American Institute of Aeronautics and Astronautics

II. Test Description


A. Facility Setup.
The McKinley Climatic laboratory at Eglin Air Force Base was chosen because it can easily generate the needed
conditions and is large enough to accommodate the test. A full-scale test article was desired because ice accretion
on propeller blades is a function of centrifugal force. In addition, recent upgrades were done that included an open
loop 12-foot-diameter reconfigurable icing tunnel that can be temporarily installed in the middle of the main
environmental chamber. See Fig. 2. The tunnel was configured with seven 300 horse power (hp), 7-foot-diameter
fans, feeding into a plenum with a 12-foot-diameter output. The air then passed through a spray bar array that
introduced water droplets into the stream.

Figure 2. The 12-foot-diameter reconfigurable icing tunnel.


At the temperatures and speeds being tested, the droplets needed to flow about 32 feet beyond the spray bar
before they became supercooled, so a 105-inch-diameter, 24-foot-long containment duct was used to prevent the
cloud from expanding in physical dimensions until the last 8 feet of travel. The engine was mounted on a thrust
table. See Fig. 3. The propeller was positioned approximately 32 feet downstream of the nozzles and approximately
8 feet from the discharge end of the containment duct. Another large duct was installed behind the propeller to duct
the engine exhaust as well as a majority of the icing cloud out of the test chamber.
The thrust stand was calibrated to a maximum of 2000 lb. The static thrust measurement accuracy for the thrust
table/load cell configuration was estimated to be approximately less than 1% of full load; however, due to engine
vibration, propeller vibration, wind speed fluctuations, electronic noise, and other factors, thrust measurement values
appeared to rapidly and continually fluctuate by as much as 5% or more. Consequently, all thrust measurements,
which were displayed and recorded for the test, were running time averages of approximately 10 to 15 seconds.

3
American Institute of Aeronautics and Astronautics

Figure 3. Engine, test stand (blue), and thrust stand (yellow).


B. Instrument and Data Acquisition
During each test run, the engine RPM, engine torque, propeller blade angle, boot voltage(s), thrust, wind
(on/off), and temperature were measured.
The engine RPM was measured from an electric pulse signal generated from the propeller synchronizer output.
This system generates one pulse for every revolution of the engine. The engine torque was measured using the
engines standard torque output signal. The beta angle on the propellers was measured real time using string pots
mounted in the propeller hub. See Fig. 4. The data from the string pots was transmitted via a wireless transmitter
mounted on the forward spinner backing plate. This system was calibrated for the full range of propeller blade
travel and was found to be accurate to within 0.5 throughout its range. This calibration took into consideration the
test temperatures and the catenary curve on the string pot wire due to the rotation speed of the engine.

Figure 4. Blade angle measurement equipment. Photo by AeroTEC.


4
American Institute of Aeronautics and Astronautics

C. Icing Conditions
Target liquid water content (LWC) and median volumetric diameter (MVD) could not be measured during the
test runs. The LWC and MVD were achieved by setting, monitoring, and recording water flow rate and atomizing
air pressure on the spray system. Calibration runs were performed before the test to determine proper water flow
and air pressure combinations needed to achieve the desired LWC and MVD required for each test point. Wind
speed could not be measured during the test, therefore, it was based on the pretest calibration of the fan speeds in the
wind tunnel. Chamber static temperature and humidity were monitored and recorded. The velocity profile along the
radius of the open jet core was measured. There was some reduction after about 75%, radius but it was determined
to be acceptable for the test. A grid was used to show the cloud spray was uniform.
D. Imaging
After each test run, the resultant ice shapes were photographed using a 13.5-Mega-pixel digital camera. Highspeed stop action video was taken during the test runs. Two high-speed cameras were used; one recorded images of
an entire propeller blade while the other recorded close-up images of the boot area of one propeller blade. Both of
the cameras were housed in enclosures to prevent ice accumulation and minimize wind vibration. The camera
synchronizing was done by triggering the cameras with the standard pulse signal output normally used for propeller
synchronization. The propeller blades were illuminated using four Arri 1200-watt compact (Hydrargyrum mediumarc iodide) HMI lamps. See Fig. 5 for camera and light placement.

Figure 5. High-speed camera and lighting placement.


E. Tracings
Tracings of the ice accretions were also taken at the conclusion of most runs. This was accomplished by
inserting a hot ice knife into the ice accretion to melt a chord-wise slot. A tracing template was then inserted into the
slot until full contact was made with the leading edge of the propeller. The ice shape was then recorded on the
tracing template using a pencil to trace the adjacent ice contour. Tracings were taken at deicing boot midspan, 50%
blade radius, and 75% blade radius.

5
American Institute of Aeronautics and Astronautics

F. Test Articles
During the test, we had planned to use five different propellers on two different engines. These test articles are
listed in Table 1.
Table 1. Planned test articles
Maximum
Engine Prop RPM
TIO-5402500
J2BD

TPE33110-511M

1591

Propeller
Blades
2aluminum
3composite
4composite
4aluminum

Propeller
Manufacturer
Hartzell
Propeller
Hartzell
Propeller
MT Propeller

Hub
HC-I2YR-1BF

Blade
F8074

Deicing
Schedule
90/90

HC-I3YR-1E

7890K

90/90

MTV-14-B

195-30a

Hartzell
Propeller

HC-B4TN5( )L

LT10282NSB5.3R

MT Propeller

MTV-27-1-E-CF-R(G)

CFRL250-55

90/90;
continuous
34/34/68;
10/60;20/60;
90/90
continuous

5composite

Due to delays in the setup, the amount of time allotted for testing was reduced from an original 12 days to 7
days. As a result, the reciprocating engine (TIO-540-J2BD) and associated propellers were not tested, and ice
accretion, as a function of RPM, could not be evaluated. All the test configurations of the TPE331-10-511M
turboporop engine were accomplished with the exception of those with 10/60 and 20/60 deicing schedules.
The five-blade composite MT propeller was a new propeller. The Hartzell four-blade aluminum propeller was
made up of both new and used blades to evaluate the effects of surface roughness on ice adhesion due to normal
blade erosion in service. Blades in the same condition, new or service condition, were opposite one another. The
blades were numbered for identification. Blades #2 and #4 were new blades, and blades #1 and #3 were service
condition blades. Blade #1 had light paint erosion at the extreme leading edge with no other wear. Blade #3 had
slightly more erosion than blade #1. Both service condition blades were considered to be in above average condition
for mid-time blades (1,468 hours into a 3,000 flight-hour overhaul interval) See Fig. 6.

Figure 6. In-service blades from Hartzell four-blade propeller.


Table 2 shows the deicing boots that were installed on the propeller assemblies listed in the order they were
tested.

6
American Institute of Aeronautics and Astronautics

Table 2. Deicing boots installed on propeller assemblies


Propeller
Assembly
Hartzell
MT Propeller
Hartzell

Deice Boot
P/N
Goodrich
4E1188-7
MT
TBD
Goodrich
4E2837-10

Deice timing
(sec)
34/34/68

Deice Schedule

Continuous

Inboard zone/outboard
zone
All blades

90/90

Opposing blades

The first test article represented the configuration of the Mitsubishi MU-2B airplane. The second configuration
did not represent any specific aircraft. The third test represents the certificated configuration of several airplanes
currently in service.
Deicing power was supplied by a 28-Volt direct current (Vdc) power supply. The power to the boots was turned
on and off using computer-controlled relays so that various timing sequences could be used during the test. The
wiring, brush blocks, slip ring, and relays were standard equipment found on an MU-2B. The MT propeller had a
slip ring that mated with the existing brush blocks. Voltage was measured between the relays and the brush blocks
to confirm that the boot was not only on but was also getting the correct voltage.
G. Test Matrix
The test setup lasted several days longer than had been scheduled. As a result, the test matrix had to be
substantially truncated and revised. Consequently, some objectives, such as icing as a function of RPM and
evaluation of scaling, were not met. A summary of the truncated test matrix is shown in Table 3.
Table 3. Test matrix used during test

RUN
No.
16
15
21
18
15A
15B
21A
3A
3B
4A
21B
19A
21C
4B
4C
26A
26B
26C
25
24
24A
24B

LWC
(g/m3)
1.04
1.04
0.36
2.44
1.04
1.04
0.36
0.33
0.33
0.57
0.36
0.1
0.36
0.45
0.52
0.4
0.52
0.36
0.36
0.66
0.4
0.36

MVD
(m)
22
22
96
15
22
22
96
16.5
16.5
16.5
96
40
96
20
20
20
20
96
96
15
20
96

OAT
(F)
12
12
12
12
12
12
12
4.6
4.6
15.2
24
12
12
12
22
12
22
24
24
19
12
12

RPM
1480
1480
1450
1520
1520
1600
1600
1450
1480
1520
1540
1535
1540
1560
1520
1485
1520
1520
1600
1596
1595
1530

TORQUE
(%)
23
30
23
21
33
38
45
39
34
42
36
35
39
36
33
40
42
42
50
42
45
36

BETA
ANGLE
()
24.5
25
25.1
24.6
26.6
26
26.2
26.2
26.2
26.2
26.2
26.2
26.2
26.2
26.7
15.2
15.2
15.2
26.2
26.2
26.2
26.1

7
American Institute of Aeronautics and Astronautics

DEICE
TIMER
34/34/68
34/34/68
34/34/68
34/34/68
NO DATA
34/34/68
34/34/68
34/34/68
34/34/68
34/34/68
34/34/68
34/34/68
34/34/68
34/34/68
34/34/68
Continuous
Continuous
Continuous
90/90
90/90
90/90
90/90

TIME
(Min.)
11.75
14.17
11.5
0.65
14
15.53
12.42
10.58
4.25
11.88
11.98
15.02
13.57
14.57
13.9
14.07
14.58
13.6
13.62
15.02
17.33
11.23

III. Discussion of Technical Issues and Challenges


A. Velocity and Pressure Limitations
The tunnel could only achieve 100 knots true air speed (KTAS) at sea level; therefore, several types of scaling
were needed so that the size and shape of the ice obtained during our test would be the same as those found at the
reference conditions which are much higher and faster. Another factor that had to be addressed was that the local
blade angle of attack (AOA) was also a function of tunnel velocity.
The planned test included both unprotected and protected runs. The unprotected runs were to be conducted with
the deicing system off and the blade radius of interest at the 75% span. The protected runs were to be conducted
with deicing on with the radius of interest at the mid-span of the deicing boot. A different scaling methodology was
used for each type of run.
For the unprotected runs, the scaling calculations used the methodology developed for 2D airfoils at NASA
Glenn by David Anderson. The objective is to find scale test conditions to simulate the reference test, such that scale
ice accretion is the same size and shape as would be obtained with the reference test. Calculations are at the
stagnation point. Essentially, MVD is found by matching the collection efficiency 0 , temperature is found by
matching the freezing fraction n 0 , and test time is found by matching the accumulation parameter Ac. Due to the
facility limitation, velocity cannot be properly scaled, but it is believed that the effect on ice shape should be small
for most conditions. However, neither centrifugal nor aerodynamic shedding is included in the analysis, and this is a
significant limitation of the methodology. There is no data to evaluate the scaling method for propeller icing, so its
validity is unknown.
The scaling methodology was applied at 75% blade. At this location, the radial speed is very high, with the
result that the 0 exceeds .90 for all but very small MVDs, and only a very small change in MVD from the reference
value was necessary to scale 0 . Similarly, only small changes in static temperature and test time were necessary for
most reference test conditions.
Andy Broeren of the University of Illinois at Urbana-Champaign performed the thermal scaling for the protected
surfaces of the propeller and recommended several scaling solutions to match reference conditions. During the test,
two of the recommended thermal scaling methods were used for comparison. The first method was a two-step
process that required raising the MVD to match the modified droplet inertia parameter K 0, and increasing LWC to
match the water catch rate. The second method was essentially the same as the first but had a third step that required
increasing the temperature to match the reference stagnation point freezing fraction.
Brian Meyer from Hartzell and Martin Albrecht from MT Propeller performed scaling so the propeller blade
AOA at the scaled conditions matched the AOA for the reference condition. Due to the twist in the propeller, one
angle would not work for the entire span of the propeller. As a result, blade angle calculations were performed at
three stations on each propeller. These stations were located at the boot center, mid-span, and span, as measured
from the center of rotation. To do this, calculations were done to determine the AOA at each of these points during
the reference condition, then more calculations were done to determine the beta angles required to produce the same
AOA at the same RPM at test airspeed. These beta angles were set at the beginning of each test run then were
allowed to change on their own due to the ice buildup.
Since testing was delayed due to setup, it was recognized that many planned runs could not be accomplished.
Because of this, it was decided early in the test to perform all runs with the propeller deicer on. The reasons for this
are as follows. First, thrust decrement due to ice accretion was a primary data output, and propeller deice-on
represented an operational configuration. Second, the primary objective of the propeller deice-off runs was to
document ice accretion along the span as a function of RPM, freezing fraction, and other parameters. The results
from the initial runs showed that this could be done with the propeller deicer on, since ice accreted both on and
outboard of the deicer boots during these runs. Some of the conditions that had been planned for deicer-off were run
with deicer-on.
Test runs that were originally planned for propeller deicer-on were accomplished at the reference conditions
rather than the calculated test conditions in the test matrix. This was done because the boot mid-span was the
reference radius for these runs. Since the 75% radius was critical for determining thrust loss, the reference
conditions were used because scaling showed the calculated test conditions to be very close to the reference
conditions at 75% radius.

8
American Institute of Aeronautics and Astronautics

B. Drag Due to Test Stand


Attempts were made to quantify the drag due to the ice accretion on the engine stand. After one run, the ice on
the stand and nacelle were cleaned off, but the propeller ice was not disturbed. The engine was restarted, the wind
turned on, and thrust measurements were taken again. These thrust measurements were then compared to the thrust
values obtained prior to deicing the test stand. The icing conditions were the SLD conditions.
Another series of tests were performed comparing a clean engine stand and nacelle to an iced engine stand and
nacelle with the engine off and the propeller feathered. These tests showed that the stand and nacelle had 300 lb of
drag clean and an additional 30 lb of drag with ice accreted during a 15-minute exposure to moderate icing
conditions. See Fig. 7. Hartzell Propeller Inc. used this data to calculate the effect of ice on the engine stand with
the propeller generating thrust, i.e., the scrub drag due to the stand being subject to higher local velocity in the
propeller wash.

Figure 7. Stand ice used for drag calculations (propeller deiced).

IV. Test Results


The full-scale propeller icing test documented in this paper provides a unique and extensive data set for the study
of propeller icing. In addition to the numerical data, the exceptional imaging data obtained makes a valuable
contribution to the documentation of propeller icing. Documentation of all data recorded during this test can be
found in the Propeller Icing test data report4.
A. Blade Material
The reciprocating engine tests, which were not accomplished, were designed to include an evaluation of different
blade material. Some limited observation of the effect of blade material on ice adhesion was made. Three test runs
were conducted with the five-blade composite MT propeller; these were runs 26A, 26B, and 26C. Run 26A and
26B were repeats of Appendix C runs 4B and 4C, respectively, and run 26C was a repeat of SLD run 21B. The
reduction in thrust at 100% torque compared well for run 26A and 26C, but the metallic thrust loss was much lower
in run 4C. See Table 4.

9
American Institute of Aeronautics and Astronautics

Table 4. Reduction in thrust composite versus metallic


Composite
Run number
26A
26B
26C

Composite
Thrust Reduction
10.0%
10.5%
10.8%

Corresponding
Metallic run
4B
4C
21B

Metallic
Thrust reduction
7.5%
2.26%
11.3%

The resultant ice shapes were similar between the two blade materials. See Figures 8-10. However, on run 26B,
the final ice shape was thicker and extended further to the tip than on run 4C, which accounts for the greater thrust
loss. One possible explanation for this difference is that the LWC did not match between the two runs. Run 26B
had an LWC of 0.52 g/m3, while run 4C had a LWC of 0.45 g/m3.
The stop action video data showed that, in all three cases, the extent of icing and shed rates were similar between
the composite and metallic propeller. Run 26B appeared to be further into a shed cycle than Run 4C, which could
also help account for the final ice shape differences and the greater reduction in thrust.
As a rule, ice shedding frequency seemed to be independent of the propeller and deicers tested, and averaged a
shed every 3-4 minutes on all test runs. Post flight inspection revealed that the blades did not all shed at once, and a
blade shed event was not apparent on the real-time thrust data.

Figure 8. Run 26A (composite, lower) vs. Run 4B (metallic, upper).

10
American Institute of Aeronautics and Astronautics

Figure 9. Run 26B (composite, lower) vs. Run 4c (metallic, upper).

Figure 10. Run 26C (composite, lower) vs. Run 21B (metallic, upper).

B. Blade Condition
There was no noticeable difference between the ice shape location and shedding frequency due to blade
condition. Figure. 11 shows the similarities in ice shapes on a typical run. The blade numbers are written on the
tips of the propeller with either an O to indicate an in service blade or an N to indicate a new blade.

11
American Institute of Aeronautics and Astronautics

Figure 11. Comparison of old vs. new blade on a typical test run.

C. Runback Ice Accretions


Runback ice accretions and resulting 20% efficiency losses, as documented in reference 2, were not observed.
This may be that the critical temperature was not tested. Not enough tests were conducted at warmer temperatures
to evaluate runback ice. The effect of blade material and RPM was not evaluated since the reciprocating engine was
not tested. There were some effects observed due to deicing timing. The larger deicing ON time did improve
inboard radius ice shedding at an intermediate ambient temperature of 12F, on the propeller in which two deicing
boots were evaluated. For the shorter ON time, ice accretion existed on the entire boot, even at the stagnation.
The continuous heating scheme provided the largest icing efficiency losses; it is not known if this is due to runback
ice.

D. SLD Ice Accretions


The SLD condition determined during calibration was 96 MVD and 0.36 LWC. This was an attempt to simulate
freezing drizzle with MVD>40, as defined in reference 5. Two temperatures were chosen: 12F to simulate the MU2B encounter, and 24F arbitrarily chosen to simulate a warm SLD condition. Thrust penalties in the SLD condition
12
American Institute of Aeronautics and Astronautics

were higher than the Appendix C icing conditions tested. The ice accretion and estimated propeller efficiency loss
of the MU-2B flight test event were approximately duplicated. See Fig. 12.

Figure 12. SLD test run compared to MU-2B.

E. Thrust
Thrust values were recorded continuously. The propeller thrust was comparable to the drag of the test stand,
while at test condition, as a result, the measured thrust was low. On several of the early runs, this resulted in the
measured thrust reducing to zero midway through the run. To overcome this limitation, the engine was run up to
100% torque to measure thrust values just before each test began and again immediately after each test. The thrust
reductions shown in Table 5 were determined by comparing the thrust measurements taken at 100% torque. The
thrust reductions for the SLD runs averaged 13.4% and had a maximum of 21.2%. The thrust reductions for
Appendix C conditions averaged 5.9% and had a maximum of 13.4%.
Table 5. Thrust reductions at maximum power using Hartzell stand drag correction
Deicer
34/34/68
Continuous
90/90

SLD 12F
21.2%
16.6%

SLD 24F
11.3%
10.8%
9.0%

Appendix C
13.4%
10.5%
5.6%

A nominal thrust penalty on the order of 10% for Appendix C icing certification is proposed, unless another
value can be substantiated. Earlier independent flight test results3,4 suggest a value of 10% would cover most
Appendix C icing encounters.

V. Conclusion
The full-scale propeller icing test documented in this paper provides a unique and extensive data set for the study
of propeller icing. In addition to the numerical data, the exceptional imaging data obtained makes a valuable
contribution to the documentation of propeller icing. Detailed test data will be published in an FAA report6.
The test results support the assumption of a nominal thrust penalty on the order of 10% for icing certification.
Another important result was that by simulating supercooled large drop (SLD) conditions, it was possible to
approximately match a nearly full blade span photographed during flight test in suspected SLD conditions.
Additional research should be accomplished to complete the objectives originally planned and to develop
analytical tools that will allow propeller performance prediction for a given icing condition and deicing design.
These would include:

13
American Institute of Aeronautics and Astronautics

Conduct additional propeller icing tests to complete the original test objectives. Evaluation of runback ice
accretions may be combined with research on other applications that use thermal deicing systems, such as
rotorcraft blades and fixed airfoils.

Conduct additional tests to measure thrust stand drag due to ice in Appendix C conditions.

Validate the spanwise accretion prediction of existing analytical methods.

Empirically measure lift and drag of propeller sections with ice shapes, simulating the accretions observed
in this test, and calculate propeller efficiency losses with the measured lift and drag.

Acknowledgments
Several organizations and individuals contributed to the planning, implementation, or reporting of the test.
Without their dedicated efforts, the value of the test program would not have been fully realized. Not all can be
recognized here, but the authors would like to acknowledge the contributions of Brian Meyer and his colleagues
from Hartzell Propeller, Gerd Muehlbauer and his colleagues from MT Propeller, Mr. Alan Farhner and his
colleagues from Goodrich, Dwayne Bell and his outstanding team at McKinley Climatic Laboratory, Lee Human
and his colleagues at AeroTEC, Mark James and his colleagues from Intercontinental Jet, Mr. Vincent Reich and
his colleagues from RSIS, Dr. David Anderson of the Ohio Aerospace Institute, and Dr. Andy Broeren of the
University of Illinois at Urbana-Champaign.

References
1

Timmons, L., Icing Investigations and Product Development on MU-2B Airplanes, SAE 2003-01-2088, presented at FAA
In-Flight Icing/ Ground De-icing International Conference, Chicago, IL, June 16-20, 2003.
2
Rodling, S., Experience From a Propeller Icing Certification, presentation at the SAE Aircraft Icing Committee meeting
in Zurich, Switzerland, September 18-20, 1989
3
Preston, G. M. and Blackman, C.C., Effects of Ice Formations on Airplane Performance in Level Cruising Flight, NACA
TN 1598, May 1948.
4
Broeren, Neel, C. B., Jr. and Bright, L. G., The Effect of Ice Formations on Propeller Performance, NACA TN 2212,
October 1950.
5
Ice Protection Harmonization Working Group, Task 2 Working Group Report on Supercooled Large Droplet Rulemaking,
September 2005
6
Pellicano, P., Dumont, C., Smith, T., Riley, J., Bell, D., and Reinhardt, E., Data From a Full-Scale Propeller Icing Test,
FAA report, to be published.

14
American Institute of Aeronautics and Astronautics

You might also like