You are on page 1of 8

Aerospace Science and Technology 10 (2006) 9299

www.elsevier.com/locate/aescte
Design of precision balance and aerodynamic characteristic measurement
system for micro aerial vehicles
Arief Suhariyono, Jong Hyun Kim, Nam Seo Goo

,
Hoon Cheol Park, Kwang Joon Yoon
National Research Laboratory for Active Structures and Materials, Articial Muscle Research Center, Department of Aerospace Engineering,
Konkuk University, 1 Hwayang-dong, Kwangjin-gu, Seoul 143-701, South Korea
Received 11 July 2005; received in revised form 22 September 2005; accepted 6 October 2005
Available online 1 December 2005
Abstract
This paper presents the design and validation of a measurement system for the aerodynamic characteristics of micro aerial vehicles. Micro
aerial vehicles are very small, have a wingspan of about 100150 mm, and operate at relatively low Reynolds numbers. A precision aerodynamic
balance was designed in order to measure the lift, drag, rolling-moment, and pitching-moment of micro air vehicles, as these parameters are too
small to be determined by general-purpose aerodynamic measurement systems. During the design process, the aerodynamic balance was analyzed
with the nite-element method in order to produce an optimal design and was calibrated as well. The calibration results and the nite element
analysis results were compared and found to be in good agreement and the nal design was determined to be acceptable. In addition, a computer-
based data acquisition system has been developed to acquire measurement data from the aerodynamic balance. Wind tunnel tests of a 2D airfoil
were performed to validate the aerodynamic measurement system. Results of the wind tunnel tests were compared with references and found to
be in good agreement. Finally, we measured aerodynamic characteristics of MAV wing model used in Batwing with the developed aerodynamic
measurement system.
2005 Elsevier SAS. All rights reserved.
Keywords: Aerodynamic measurement; Wind tunnel test; Aerodynamic balance; Micro aerial vehicle (MAV); Low Reynolds number
1. Introduction
Recently, development of small aircrafts is gaining wider
attention due to their potential applications. These aircrafts,
called micro aerial vehicles (MAVs), were initially designed
for military surveillance missions. MAVs have a wingspan
of about 100150 mm and a total mass of about less than
100 grams. They are also capable of operating at speed of less
than 70 km/hr. These specications make MAVs operate at a
low Reynolds number of below 200,000. Fig. 1 shows the typ-
ical relation between mass and Reynolds number of ying ob-
jects including conventional aircrafts and MAVs. In addition, a
video camera can be mounted on the MAV to transmit real-time
images of nearby objects, which helps the operator to control
*
Corresponding author. Tel.: +82-2-450-4133; fax: +82-2-444-6670.
E-mail address: nsgoo@konkuk.ac.kr (N.S. Goo).
these vehicles remotely. This is possible because the camera
and sensor can now be miniaturized.
Besides surveillance missions, combat missions consti-
tute a potential area for future MAV applications as they
are capable of carrying specic martial arms such as small
bombs. MAVs are expected to y for 30 minutes or even
for one hour in order to perform special missions. For this
purpose, the development of a long-lasting battery is essen-
tial. Eventually, it is expected that MAVs will be controlled
autonomously using a smart ight control system. In order
to fulll this requirement, the development of a new-class
MAV with greater payload capacity is necessary so that the
MAV can carry the electronic systems needed for autonomous
ight.
In 2005, a MAV called Batwing was developed by the au-
thors and won second prize in the endurance division of the 7th
international MAV competition held in Florida, USA and rst
prize in the same division of the 9th international MAV compe-
1270-9638/$ see front matter 2005 Elsevier SAS. All rights reserved.
doi:10.1016/j.ast.2005.10.004
A. Suhariyono et al. / Aerospace Science and Technology 10 (2006) 9299 93
Fig. 1. Reynolds number range of MAVs [5,6].
Fig. 2. Micro aerial vehicle of Konkuk University (Batwing).
tition held in Seoul, Korea. The Batwing, shown in Fig. 2, has
a wingspan of 130 mm and a mounted video camera system,
and can y for 15 minutes. In order to improve its endurance
and increase payload, studies on MAV performance including
aerodynamics, ight performance, control, and dynamic stabil-
ity are essential.
Studies of MAVs and their low Reynolds number charac-
teristics have recently been performed and published [2,3,57].
However, there is still not sufcient data to innovatively im-
prove the performance of MAVs. In this regard, important aero-
dynamic characteristics can be experimentally obtained through
wind tunnel tests. However, a precision measurement system to
accurately acquire aerodynamic data is needed.
This paper presents the design of an aerodynamic charac-
teristic measurement system for MAVs. First, a precision bal-
ance design is introduced and a nite element analysis of the
balance is presented. A data acquisition system including the
wind tunnel apparatus and its verication are then delineated.
The verication is essential to assess whether the measure-
ment system is suitable for measuring the aerodynamic per-
formance of MAVs. The verication was performed with a
2D low-Reynolds-number airfoil, as in Ref. [5]. As a practical
application of the present aerodynamic characteristic measure-
ment system, we measured lift and drag of the Batwing MAV
wing model.
2. Design of balance
2.1. Overview
An aerodynamic balance was designed to accommodate all
MAV aerodynamic tests including MAVs 3D wings and full-
scale models. As noted in the Introduction, MAVs y at a low
Reynolds number regime. The balance must be designed ac-
curately and precisely because boundary layers in the wind
tunnel are very sensitive to small disturbances. Since aerody-
namic forces and moments produced by an MAV are extremely
small, a great deal of care must be exercised to obtain accurate
data.
First, the balance was designed with commercial CAD soft-
ware. In order to choose the optimal locations of all sen-
sors in the preliminary design stage, the balance was analyzed
through a nite element analysis (FEA). This approach mini-
mized time and cost consumption in the design process. The
FEA was performed using commercial nite element software,
MSC/NASTRAN [11]. Subsequently, the balance was manu-
factured, built, and calibrated.
2.2. Specication
The balance is composed of four sensors: a lift sensor, a drag
sensor, a pitching-moment sensor, and a rolling-moment sen-
sor. In addition, the balance can produce angle-of-attack (AOA)
changes from 10 to 40 degrees manually with the smallest
increment of 2 degrees. Fig. 3 shows the nal design of the
aluminum aerodynamic balance including the locations of all
sensors. It has dimension of 88 mm15 mm318 mm. All
sensors were constructed by strain gages in a Wheatstone full-
bridge conguration because the measured forces and moments
are extremely small. An excitation voltage of 5 V or 10 V
is used for all the strain gage bridges. The strain gages for
force measurement can withstand a maximum load of 10 N,
Fig. 3. Aerodynamic balance design and locations of sensors.
94 A. Suhariyono et al. / Aerospace Science and Technology 10 (2006) 9299
Fig. 4. Finite element model of the balance.
whereas the strain-gages for moment measurement can endure
a maximum load of 50 Nmm. Furthermore, all sensors have
sensitivity of 0.1% of maximum load.
2.3. Finite element analysis (FEA)
Performance of a balance is affected by two criteria: shape
of the balance and locations of the sensors. For the balance to
be effective, each sensor should be able to capture either aero-
dynamic force or moment without errors. Errors can arise from
coupling between sensors and external disturbances such as air-
ow, windshield, and wind tunnel interaction. A nite element
analysis was performed to reduce possible errors.
Fig. 4 displays the nite element model of the balance.
Fixed boundary conditions were applied on the bottom side
to describe the actual boundary conditions. Loads were ap-
plied at points on the top of the balance model, as shown
in Fig. 4. During the analysis, nite element models were
made with 4,000 TETR4 nite elements after performing an h-
convergence test [1]. The FEA results and calibration data were
compared to verify that the FEA results were reliable.
2.4. Validation of balance design
Design and manufacture of the balance based on the nite el-
ement analysis results were justied by a calibration procedure.
Good agreement between the experimental and nite element
analysis results would indicate that the FEA assisted balance
design is useful and effective. The setup of the calibration ap-
paratus is shown in Fig. 5: it includes a calibration base, a roller
assembly with supporting posts, several calibration weights, a
switch and balance unit, and a strain indicator.
In order to calibrate the balance, we rst applied force or
moment, and then measured strains of all sensors shown in
Fig. 3. The post on the left-hand side is used for pitching-
moment calibration while the post behind the balance is used
Fig. 5. Calibration setup and apparatus.
for rolling-moment calibration. In Fig. 5, signals from all sen-
sors are delivered to the switch and balance unit in order to
adjust and select signals from the sensors. The signal is then
sent to the strain indicator.
A comparison of strain obtained by FEA and calibration for
each load on all sensors is presented in Table 1 together with
the relative error in percentage. The relative error was calcu-
lated based on the main diagonal strain of the FEA results as
expressed in Eq. (1).
E
c
=
S
S
main,FEA
100%, (1)
where E
c
is the relative error, S is the difference between the
FEA strain and the measured strain, and S
main,FEA
dened as
the main diagonal strain is the corresponding strain component
for the applied load in the FEA results.
As shown in Table 1, the calibration results agree well with
the FEA results. The relative errors were less than 4%, ex-
cept 12% relative error of the rolling-moment sensor. The cause
of the larger relative error of the rolling-moment sensor might
be complexity of geometry. Because the rolling-moment is not
very important in the aerodynamic study of an MAV, except
in studying the MAVs dynamic performance of the control
surface, which requires a six degrees of freedom balance (6-
DOF balance), the rolling-moment sensor could be used to
check the rolling stability of MAV models qualitatively. Con-
sequently, it could be concluded that the balance design based
on the FEA could be regarded as a reasonable and acceptable
approach.
3. Measurement system apparatus
3.1. Description
The measurement system consists of four components: a
subsonic wind tunnel system, the aerodynamic balance, a data
acquisition (DAQ) system, and data processing software. The
data acquisition system has many components including ampli-
ers, a terminal block, and an analog digital converter (ADC)
card and DAQ software within a computer. Data processing
software was developed for processing all obtained data from
A. Suhariyono et al. / Aerospace Science and Technology 10 (2006) 9299 95
Table 1
Comparison of strain obtained by FEA and calibration for each load on all sensors
Load (gr.) Load type
strain
from FEA
strain
from calibration Relative error (%)
Sensor Sensor
Drag Lift Roll Pitch Drag Lift Roll Pitch Drag Lift Roll Pitch
0 0 0 0 0 0 0 0 0
50 Drag 91 2 0 2 93 0 1 2 2.20 2.20 1.10 0
100 182 3 0 4 184 0 2 4 1.10 1.65 1.10 0
0 0 0 0 0 0 0 0 0
50 Lift 1 93 0 0 0 92 0 0 1.08 1.08 0 0
100 2 186 0 0 0 185 0 1 1.08 0.54 0 0.54
0
Rolling
Moment
0 0 0 0 0 0 0 0
25 0 0 25 0 0 0 22 0 0 0 12.00 0
50 0 0 50 1 0 0 44 0 0 0 12.00 2.00
0
Pitching
Moment
0 0 0 0 0 0 0 0
25 0 0 0 28 1 0 1 29 3.57 0 3.57 3.57
50 0 0 0 56 2 0 1 58 3.57 0 1.79 3.57
Fig. 6. Schematic of the low-speed wind tunnel.
the DAQ system. Thus, the aerodynamic characteristics data
from the balance can be displayed graphically.
3.2. Wind tunnel system
The wind tunnel system includes a wind tunnel with a test
section, a motor together with a fan, a controller, and a velocity
measurement device. Fig. 6 schematically illustrates the wind
tunnel system which is open circuit type for low speed tests.
The fan, driven by a variable speed electric motor, provides air-
ow with a velocity from 0 to 30 m/s.
As shown in Fig. 6, the wind tunnel is composed of a fan
duct, a diffuser, an anti-turbulence section, a contraction sec-
tion, and a test section. The diffuser behind the fan duct decel-
erates the air and also gradually transforms the circle contour to
a square. Honeycomb anti-turbulence screens give small free-
stream turbulence intensity to the downstream sections includ-
ing the test section. The contraction section provides a contrac-
tion ratio 9 : 1 and a low turbulence level of 0.3% in the test
section. The cross-section of the test section is 30 cm by 30 cm
square with a transparent wall. Whereas the balance beside the
windshield and the 2-D wing model provide the blockage ratio
of 510% for all AOA range of the experiment, the maximum
blockage ratio is 3.35.3% for wing model of Batwing MAV.
The maximum blockage ratio is about 5% when attaching full
model of Batwing MAV with 150 mm wing span.
At the bottom side, the test section also has a bearing hole
in which a windshield was placed to enclose the balance. The
Fig. 7. Equipment around the test section.
bearing hole was designed to produce side slip-angle changes
of 90 degrees to turn left and right direction. The windshield
was designed to cover the strut of the balance from distur-
bance by the free air stream and to preserve smooth airow
around the model. The ow speed was measured directly by
digital micro-manometer through pitot tube mechanism. Fig. 7
shows the bearing hole, the windshield, and a digital velocity-
measurement device.
3.3. Data acquisition system (DAQ)
A data acquisition system was designed to acquire data from
the sensors easily and process data automatically. The data ac-
quisition system can acquire signals from all sensors of the
balance simultaneously. The data acquisition system consists
of strain-gages in a full-bridge circuit, ampliers, noise lters,
a terminal block, an analog-to-digital converter (ADC) and PC-
based graphical programming software. Fig. 8 schematically
displays the data acquisition system.
96 A. Suhariyono et al. / Aerospace Science and Technology 10 (2006) 9299
Fig. 8. Data acquisition system.
Fig. 9. Lift calibration graph for the data acquisition system.
The signal conditioning strain ampliers magnify the volt-
age signal of the bridge from 1 to 1000 times. The ampliers
also include four low-pass active Butterworth lters from fre-
quency 10 Hz to 10 kHz. A terminal block is used to synchro-
nize and mix four output signal-voltages into one signal at one
time alternately. A PCI DAQ-card, or 12-bit analog-to-digital
converter, is connected to the terminal block to convert analog
signals to digital signals. All obtained data are processed using
a LabVIEW

6.0 based program. In order to reduce coupling


noise from each sensor, digital ltering in LabVIEW

was im-
plemented on the developed program.
3.4. Calibration of measurement system
After integrating the balance and the data acquisition sys-
tem, calibrations were performed to check the linearity of the
balance and the couplings of all sensors. Whereas the previous
calibration was implemented to assess the design of the bal-
ance based on the FEA results, the present calibrations were
performed in order to calibrate the complete measurement sys-
tem including the balance and the data acquisition using a
LabVIEW

based program.
Figs. 912 show calibration curves of all the sensors. The
gures indicate that each sensor is in linearity and the couplings
are very small compared to the main diagonal strain. Conse-
quently, the couplings could be ignored in the measurement of
force and moment.
Fig. 10. Drag calibration graph for the data acquisition system.
Fig. 11. Rolling-moment calibration graph for the data acquisition system.
Fig. 12. Pitching-moment calibration graph for the data acquisition system.
From calibrations, a calibration matrix converting force to
voltage was obtained. Thus, an inverse of the calibration ma-
trix, which converts voltage to force, could be determined based
on the calibration matrix. The inverse calibration matrix can be
written as follows:

D
L
M
r
M
p

1.409 0 0 0
0.079 1.409 0 0
3.134 5.634 100 0
0 0 0 50

V
d
V
l
V
r
V
p

, (2)
where D, L, M
r
, and M
p
are drag, lift, moment and voltage, re-
spectively, and the subscripts d, l, r, and p stand for drag, lift,
rolling moment, and pitching moment. Eq. (2) indicates that
A. Suhariyono et al. / Aerospace Science and Technology 10 (2006) 9299 97
there are three couplings (off-diagonal terms): the drag com-
ponent in the lift measurement, and lift and drag components
in the rolling-moments measurement. The couplings were so
small that they could be ignored.
3.5. Data processing
The obtained data from the DAQ system was processed by
a MATLAB

6.5 based program. This program was developed


to process the data, and as such it yielded nal data such as
the calibration matrix as well as aerodynamic graphs automati-
cally using output les of a LabVIEW

based program. Aero-


dynamic graphs such as lift coefcient versus AOA, moment
coefcient versus AOA, and drag polar could be obtained from
the MATLAB

based program.
3.6. Measurement uncertainty
There might be several sources of uncertainty in the data ac-
quisition system and the effect of free-stream turbulence. From
the calibration, it was veried that the performance of the bal-
ance showed linearity, repeatability, and small hysteresis.
Uncertainty in the data acquisition system was computed
using the KlineMcClintock technique [4,6]. This uncertainty
included a quantization error and uncertainty arising from the
standard deviation of a given mean output voltage. The quanti-
zation error (e
q
) is formulated as [4]:
e
q
=
1
2

Range in volts
2
M

, (3)
where M is the number of bits of the analog-digital converter.
If the gain increases, the standard deviation of the mean will
also increase, but the ratio of the standard deviation to the mean
will basically remain the same. Since the measurement system
used 12 bits of ADC and a range of voltages of 20 (10 V), the
uncertainty was small enough to be neglected.
Mueller et al. showed that an increase in free-stream turbu-
lence intensity reduces the minimum drag and eliminates some
of hysteresis in lift and drag coefcients [8]. The cross-section
of test section was divided into 15 segments, 3 horizontally by 5
vertically. The balance and model were installed in the test sec-
tion and the wind tunnel was then operated. Velocity of airow
was measured in each segment of the test section. The results
showed that the velocity of all segments were uniform, except
the velocities of segments near the wall of the test section. This
indicates that the airow was laminar around the balance and
model.
4. Wind tunnel tests
4.1. Description
Low-Reynolds-number wind tunnel tests on a 2D-airfoil
were performed to validate the measurement system. The
present wind tunnel results were compared with reference data
and numerical results. As such, performance of the present
aerodynamic measurement system could be assessed in terms
Fig. 13. SA7038 airfoil.
Fig. 14. 2D wing model.
of its capacity for accurately measuring the aerodynamic per-
formance of MAVs.
4.2. Low Reynolds number airfoil
An SA7038 airfoil, shown in Fig. 13, was used in the wind
tunnel tests for validation of the aerodynamic measurement sys-
tem. This airfoil was used for a radio-controlled (R/C) model
aircraft [5]. Experimental results of this study were compared
with those of Selig et al. [5,9,10] and veried. Besides the
reference experimental-results, the numerical results were also
presented to conrm our experimental results. The numeri-
cal results were acquired from the Hanley Innovations Wing
Analysis plus Version 1.1 [12].
A two-dimensional wing with an SA7038-airfoil cross-
section was used in the wind tunnel tests. The wing has a span
of 30 cm and a chord length of 9.5 cm. Fig. 14 shows the 2D
wing model used in the wind tunnel tests. It is made of acry-
lonatrile butadiene styrene (ABS) for most of the wing and
balsa wood for the trailing edge.
4.3. Apparatus and setup of wind tunnel tests
In the wind tunnel tests, we used the wind tunnel system,
the aerodynamic balance, the data acquisition system, and the
data processing program. The airfoil model was bolted on the
aerodynamic balance in the test section as shown in Fig. 15. The
strut of the aerodynamic balance was covered by a windshield
mounted on the bearing hole. All sensors of the balance were
connected to the data acquisition system, which was linked to
the PC-based data processing program.
The wind tunnel tests were performed at a free-stream ve-
locity of 15 m/s and a Reynolds number of 98,000. The ex-
perimental results were compared with experimental results
from reference data at Reynolds numbers of 10
5
and 149,000.
Sideslip angle was set to zero degree and the angle of attack was
set from 4 to 14 degrees with an increment of 2 degrees. The
wind tunnel tests were performed three times to obtain reliable
98 A. Suhariyono et al. / Aerospace Science and Technology 10 (2006) 9299
Fig. 15. Attaching 2D wing model on the balance in the test section.
Fig. 16. 2D lift coefcient graph of the wing with SA7038 airfoil.
Fig. 17. 2D drag coefcient graph of the wing with SA7038 airfoil.
data. The presented results are means of all acceptable results
from the wind tunnel tests.
5. Results and discussion
Figs. 1618 show C
l
versus AOA, C
d
versus AOA, and C
m
versus AOA of a wing with an SA7038 airfoil, respectively. C
l
,
C
d
and C
m
denote 2D lift coefcient, 2D drag coefcient, and
2D pitching-moment coefcient, respectively. Each graph also
presents experimental results from references at two different
Reynolds numbers [9,10] and numerical results.
Fig. 18. 2D pitching-moment coefcient graph of the wing with SA7038 airfoil.
Fig. 19. 2D rolling-moment coefcient graph of the wing with SA7038 airfoil.
In the lift coefcient graph, shown in Fig. 16, it can be seen
that the experimental results are in good agreement with the
reference results and the numerical result. For AOA larger than
12 degrees, the measured C
l
is lower than the reference data and
it is already in the stall region that the ow becomes unstable.
From the drag coefcient graph, shown in Fig. 17, it can be
observed that the measured drag coefcients are slightly smaller
than the references and the numerical result. However, each
graph has a similar slope, except in the high AOA. The slope
of the present data is higher than that of the reference graph at
AOA higher than 12 degrees.
In the pitching-moment coefcient graph, shown in Fig. 18,
it can be seen that the measured coefcient is smaller than the
reference data. The slope of the experimental graph is similar
to the reference results. However, numerical result had large
difference with experimental ones. We thought that this was due
to the limitation of numerical program.
Fig. 19 shows the measured rolling-moment from the tests.
The rolling-moment results were measured only to conrm the
quality of rolling stability. From Fig. 19, it can be observed
that the 2D airfoil model was stable at low AOA. However,
an increase in C
r
was measured near 6 degrees of AOA. This
could be attributed to the aspects that the 2D airfoil model is not
perfectly symmetrical. Subsequently, the rolling stability main-
tained a constant value of approximately 12 degrees AOA.
With the veried aerodynamic characteristic measurement
system, we measured lift and drag of an MAV wing model. The
A. Suhariyono et al. / Aerospace Science and Technology 10 (2006) 9299 99
Fig. 20. Attaching 3D MAV wing model on the balance in the test section.
Fig. 21. Drag polar of the MAV wing model.
wing model, which was used in the Batwing, was manufactured
from berglass and balsa wood. The average surface roughness
was from 1.11 to 1.64 mR
a
. It has dimension of 110 mm wing
span and 100 mm chord. The wing model was bolted on the
balance as shows in Fig. 20. The experiments were performed
for AOA changes from 4 to 12 degrees. Fig. 21 shows drag
polar of the MAV wing model for the ow speed of 10 m/s
and 12 m/s. This drag polar data can be used as fundamental
data for precise design of MAVs.
6. Conclusion
An aerodynamic characteristic measurement system was de-
signed and fabricated to investigate the aerodynamic perfor-
mances of MAVs. The system is composed of a subsonic wind
tunnel, a newly designed balance, a data acquisition system, and
a data processing program. The balance was designed through
FEA. The FEA results were compared with the calibration re-
sults and found to be in good agreement with a maximum
relative error of 3.57%, except error in the rolling-moment sen-
sor of 12%. The data acquisition system and data processing
software were developed to acquire data from the balance au-
tomatically. Calibrations and measurement uncertainty studies
indicated that the measurement system displayed qualities of
linearity, repeatability, and small hysteresis. In addition, the
measurement system had so small couplings that they could be
neglected. 2D-airfoil low-Reynolds-number wind-tunnel tests
were performed to verify the effective operation of all com-
ponents of the measurement system. Comparisons between the
experimental results and reference data showed that they were
in good agreement. The aerodynamic test results of the MAV
wing model used in the Batwing were also provided as a prac-
tical application. Based on the investigations in this study, it
could be concluded that the new aerodynamic measurement
system is acceptable for measuring the performance of full
three-dimensional MAV models as well as MAV wings.
Acknowledgements
The present work was supported by the Korea Research
Foundation Grant (KRF-2004-005-D00045, KRF-2004-005-
D00046, KRF-2004-005-D00047). The authors appreciate this
nancial support.
References
[1] K.J. Bathe, Finite Element Procedures, second ed., Prentice-Hall, Engle-
wood Cliffs, NJ, 1996.
[2] T.F. Burns, Experimental studies of the Eppler 61 and Pfenninger 048 air-
foil at low Reynolds numbers, Masters Thesis, The University of Notre
Dame, 1981.
[3] G.A. Fleming, S.M. Bartram, M.R. Waszak, L.N. Jenkins, Projection
moir interferometry measurement of micro air vehicle wings, in: Proc.
of SPIE Conf. on Optical Diagnostics for Fluids, Solids and Combustion,
No. 4448-16, 2001.
[4] S.J. Kline, F.A. McClintock, Describing uncertainties in single-sample ex-
periments, Mech. Engrg. 75 (1) (1953) 38.
[5] T.J. Mueller, Fixed and Flapping Wing Aerodynamics for Micro Air Ve-
hicle Applications, Progress in Astronautics and Aeronautics, vol. 195,
AIAA, Reston, 2001.
[6] T.J. Mueller, Aerodynamic measurements at low Reynolds numbers for
xed wing micro air vehicles, RTO documents, Notre Dame, 1999.
[7] T.J. Mueller, T.F. Burns, Experimental studies of the Eppler 61 airfoil at
low Reynolds numbers, AIAA Paper, #82-0345, 1982.
[8] T.J. Mueller, L.J. Pohlen, P.E. Conigliaro, B.J. Hansen Jr., The inuence
of free-stream disturbances on low Reynolds number airfoil experiments,
Experiments in Fluids 1 (1983) 314.
[9] M.S. Selig, J.J. Guglielmo, A.P. Broeren, P. Giguere, Summary of Low-
Speed Airfoil Data, vol. 1, SoarTech Publications, Herk Stokely, 1995.
[10] M.S. Selig, C.A. Lyon, P. Giguere, C.N. Ninham, J.J. Guglielmo, Sum-
mary of Low-Speed Airfoil Data, vol. 2, SoarTech Publications, Herk
Stokely, 1996.
[11] USERs guide manual of MSC/NASTRAN, version 2005.
[12] http://www.hanleyinnovations.com/.

You might also like