You are on page 1of 16

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/276481983

Effects of Mach Number and Specific Heat Ratio on Low-Reynolds-Number


Airfoil Flows

Article in AIAA Journal · June 2015


DOI: 10.2514/1.J053468

CITATIONS READS

37 2,199

4 authors:

Masayuki Anyoji Daiju Numata


Kyushu University Tokai University
58 PUBLICATIONS 439 CITATIONS 66 PUBLICATIONS 561 CITATIONS

SEE PROFILE SEE PROFILE

Hiroki Nagai Keisuke Asai


Tohoku University Tohoku University
224 PUBLICATIONS 1,368 CITATIONS 346 PUBLICATIONS 4,020 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Development of Mars Airplane View project

All content following this page was uploaded by Masayuki Anyoji on 16 April 2018.

The user has requested enhancement of the downloaded file.


AIAA JOURNAL
Vol. 53, No. 6, June 2015

Effects of Mach Number and Specific Heat Ratio


on Low-Reynolds-Number Airfoil Flows

Masayuki Anyoji∗
Japan Aerospace Exploration Agency, 252-5210 Sagamihara, Japan
and
Daiju Numata,† Hiroki Nagai,‡ and Keisuke Asai§
Tohoku University, 980-8579 Miyagi, Japan
DOI: 10.2514/1.J053468
The effects of Reynolds number, Mach number, and gas species (air and CO2 ) on aerodynamic characteristics of a
thin flat plate and a NACA 0012-34 airfoil were investigated under low-Reynolds-number (Re ! 0:43 × 104 to
4.1 × 104 ) and high-subsonic-flow (M ! 0.1 to 0.6) conditions. In addition to lift and drag measurements by a two-
component balance system, the pressure-sensitive paint technique was applied to measure pressure profiles on the
Downloaded by KYUSHU UNIVERSITY OOA on July 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/1.J053468

model surface. For the flat plate, the Reynolds number moderately affects the lift and drag characteristics because of a
simple behavior of the leading-edge separation bubble; the length of the separation bubble increases as the angle of
attack increases. By contrast, the Mach number and specific heat ratio contribute little to the aerodynamic
performance. For the NACA 0012-34 airfoil, the lift curves are highly dependent on the Reynolds number because of
the formation, shift, and burst of the separation bubble, whereas the compressibility affects only the stall
characteristics. The specific heat ratio has little effect on the aerodynamic performance. In common for both airfoils,
it was observed that the Mach-number effect allows for the delay of a laminar–turbulent transition and reattachment
of the separated shear layer.

Nomenclature sharp leading edge (e.g., a reversed NACA 0012 wing or a thin
Cd = drag coefficient wedge) develops more lift than the symmetrical NACA 0012 airfoil,
Cl = lift coefficient which is satisfactory for Re > 7.0 × 104. Schmitz [6,7] suggested
Cp = pressure coefficient that a sharp leading edge fixes the separation point at the edge and
I = luminescence intensity can improve its Reynolds-number dependence on aerodynamic
KSV = Stern–Volmer constant performance. By contrast, Ohtake et al. [8] studied the effects of low
P = static pressure, Pa Reynolds numbers (Re ! 1.0 × 104 to 1.0 × 105 ) on the aero-
Pt = total pressure, Pa dynamic performance of the NACA 0012. Their results indicated that
M = Mach number the nonlinear lift curves were susceptible to a change in the Reynolds
Re = Reynolds number (characteristic length equal to chord number.
length of airfoil) Meanwhile, as a more challenging aerodynamic design for a low-
α = angle of attack, deg Reynolds-number aircraft, Mars airplanes have been recently
γ = specific heat ratio proposed by NASA [9,10] and a Japanese research group [11,12] for
wide-range exploration with no dependence on the complex
topography of Mars and detail science measurements at a few
I. Introduction kilometers above the Mars surface. Contrary to the Earth’s atmo-

A ERODYNAMICS and flow physics at low Reynolds numbers


as represented by micro air vehicles [1], high-altitude, long-
endurance, unmanned aerial vehicles [2], and insect flight [3] have
sphere, the Martian atmospheric density is only approximately one-
hundredth that of the Earth. Such a low-density condition leads to a
low-Reynolds-number flight, Re ! O"104 #. Moreover, the Mars
been extensively studied. Unlike that at high Reynolds number, airplane will fly at relatively high speeds to produce sufficient lift and
airfoil performance at low Reynolds numbers has specific char- ensure a stable flight under windy atmospheric conditions, and then
acteristics such as a nonlinear lift curve caused by the formation or the flight speed will easily reach subsonic velocities because of the
burst of a laminar separation bubble. In addition, the maximum lift- low sound speed in the CO2 -based and low-temperature Martian
to-drag ratio of airfoils significantly deteriorates in the chord at atmosphere. If a propeller is used as the propulsion device, the tip
Reynolds numbers from 104 to 105 because of earlier flow separation Mach number of the blade also would reach subsonic speeds. Under
[4]. For Re < 7.0 × 104, Laitone [5] showed that an airfoil with a such a subsonic flow condition at low Reynolds numbers, it is
anticipated that the flowfield around the wing will become highly
Presented as Paper 2010-4627 at the 40th Fluid Dynamics Conference and complicated, with a strong interaction between viscous and com-
Exhibit, Chicago, IL, 28 June–1 July 2010; received 12 March 2014; revision pressibility effects. In addition, the specific heat ratio may affect
received 30 June 2014; accepted for publication 6 July 2014; published online
aerodynamic performance because the specific heat ratio differs
7 October 2014. Copyright © 2014 by Masayuki Anyoji. Published by the
American Institute of Aeronautics and Astronautics, Inc., with permission. between CO2 and air. Therefore, it is necessary to evaluate the
Copies of this paper may be made for personal or internal use, on condition similarity rule in the low-Reynolds-number and high-subsonic
that the copier pay the $10.00 per-copy fee to the Copyright Clearance Center, region; however, experimental data of airfoils are very limited in such
Inc., 222 Rosewood Drive, Danvers, MA 01923; include the code 1533-385X/ an uncommon flow region, as shown in Fig. 1 [13].
14 and $10.00 in correspondence with the CCC. A low-density wind tunnel, named the Mars Wind Tunnel (MWT),
*Project Researcher, Department of Space Flight System. Member AIAA. was developed at Tohoku University in 2007 as a test facility to

Assistant Professor, Department of Aerospace Engineering. Member evaluate the aerodynamic performance of airfoils under low-
AIAA.

Associate Professor, Department of Aerospace Engineering. Senior Reynolds-number and high-subsonic-speed conditions. Initially, air
Member AIAA. was used as the working gas, but a modification was made to allow
§
Professor, Department of Aerospace Engineering. Associate Fellow the tunnel to operate using CO2 (CO2 mode). We investigated the
AIAA. operational characteristics and the flow quality in the test section
1640
ANYOJI ET AL. 1641

under both air and CO2 modes. The flow calibration tests confirmed
that the MWT has a good flow quality, and its operational range
allows us to independently investigate each effect of the Reynolds
number, Mach number, and specific heat ratio on aerodynamic
performance by changing the total pressure, flow speed, and gas
species [14,15]. Furthermore, the pressure-sensitive paint (PSP)
measurement technique was developed, which allows us to optically
capture the pressure distribution on airfoils even under low-pressure
conditions [16,17], and its measurement system was installed into the
MWT. However, this PSP technique is limited for use in the air mode
and is not applicable in the CO2 mode at present.
In this study, we conducted wind-tunnel tests at low Reynolds
number (Re ! 0.43 × 104 to 4.1 × 104 ) and high subsonic flow
(M ! 0.1 to 0.6) in both air and CO2 modes using the MWT and
evaluated for each the effect of the Reynolds number, Mach number,
and the specific heat ratio on the aerodynamic performance of airfoils
at low Reynolds numbers. A thin flat plate was used as a standard
Fig. 2 Mars Wind Tunnel at Tohoku University.
model. Also, we selected the NACA 0012-34 as the second model
because a symmetric airfoil markedly shows the effect of the
Downloaded by KYUSHU UNIVERSITY OOA on July 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/1.J053468

Reynolds number on aerodynamic performance, and a comparatively is driven by a supersonic ejector located downstream of the test
small leading-edge radius is expected to prevent an early laminar section. The ejector consists of five pipes, with each having six
separation in the vicinity of the leading edge. The flowfield around equally spaced small orifices. The ejection of high-pressure gas from
the airfoils was analyzed on the basis of pressure distribution. these nozzles reduces adjacent pressure and induces flow in the
Eventually, we discuss the association between the change of the test section. In preliminary flow calibration tests using CO2 , we
flowfields and aerodynamic performance. confirmed that CO2 solidification does not occur during CO2 gas
ejection. Refer to our previous studies for the further specifications of
the MWT [15].
II. Mars Wind Tunnel
The MWT, as shown in Fig. 2, is mainly composed of a vacuum
chamber, an induction-type wind tunnel, and a buffer tank. The III. Measurement Systems and Techniques
vacuum chamber and buffer tank are connected by a flexible pipe.
A. Force Balance
A butterfly valve with a proportional–integral–derivative pressure
controller is inserted into the connecting pipe. The vacuum chamber The two-component balance system consists of two load cells for
is cylindrical, 5000 mm in length, and 1800 mm in inner diameter. lift and drag measurements and a stepping motor for changing the
The induction wind tunnel is placed inside the vacuum chamber. This angle of attack. The range of load cells for lift measurement (A&D
arrangement allows wind-tunnel tests to be conducted at low pressure AC4101-K006) and drag measurement (A&D AC4101-G600) is 60
and the test gas to be replaced with CO2 . The induction tunnel, which and 6 N, respectively. The uncertain accuracies of the lift load cell and
has a total length of 3490 mm, is made of aluminum alloy. One the drag load cell are $9.0 × 10−3 and $9.0 × 10−4 N, respectively.
honeycomb and five screens are present in the settling chamber to The output signals from the load cells are amplified by a dc strain
reduce the turbulence level in the test section, which has a cross amplifier (NISSHO-ELECTRIC-WORKS, DSA-100A). The angle
section of 100 × 150 mm. The upper and lower test-section walls resolution of the stepping motor is 1.44 × 10−2 deg, and the accu-
have an inclination angle of 1.3 deg in each side for correcting the racy is 3.4 × 10−2 deg.
development of the boundary layer. Thus, the cross section at the test-
section center is 159.1 × 150 mm. The blockage rate of the model B. Pressure-Sensitive Paint
frontal area to the test-section cross-sectional area is approximately 1. Theory
5.5% at α ! 10 deg but is 8.1% at α ! 15 deg at the test section PSP is a coating-type sensor consisting of luminescent molecules
center when the test model with 50 mm chord length is used. Barlow and a binder. When illuminated with light at an appropriate wave-
et al. [18] suggested that the maximum ratio of the blockage rate of length, the sensor molecules in PSP are elevated to the excited state.
7.5% should be used unless errors of several percent can be accepted. The excited molecules return to the ground state through several
Hence, the aerodynamics measurement error due to the blockage is photochemical mechanisms: luminescence, thermal deactivation,
comparatively small at low angles of attack less than α ! 10 deg, but and oxygen quenching. The principle of PSP is based on oxygen
our results may include errors of several percent. We consider that quenching. In the presence of oxygen molecules, the energy of
these several percent errors at high angles of attack do not affect our excited molecules is transferred to oxygen molecules, and no lumi-
discussion. The turbulence intensity is less than 0.84%. To achieve nescence is emitted. As a result, the luminescence intensity decreases
high subsonic speeds under low-pressure conditions, the wind tunnel with increasing oxygen concentration. Theoretically, the relationship
of the luminescence intensity I and the oxygen concentration %O2 &
is expressed by the following relation, known as the Stern–Volmer
relation:

I0
! 1 ' K SV %O2 & (1)
I

where KSV is the Stern–Volmer constant, and the subscript “0”


represents the vacuum condition.
For air, the pressure is proportional to the oxygen partial pressure
(PO2 ! 0.21 × P for air). Therefore, the Stern–Volmer relation can
be expressed by the pressure p and luminescence intensity I:

I ref p
! A"T# ' B"T# (2)
Fig. 1 Martian atmospheric flight (created referring to [6]). I pref
1642 ANYOJI ET AL.

where the subscript “ref” represents the reference condition, and IV. Test Model and Experimental Condition
A"T# and B"T# are calibration coefficients. In wind-tunnel tests, the A. Test Model
surface pressure can be calculated from the ratio of luminescence Figure 4 shows the airfoils of the test models. The chord length of
intensity images obtained under wind-on and wind-off (reference) 50 mm and span of 100 mm are common. A flat plate of 5% thickness
conditions. with a blunt leading edge and the NACA 0012-34 were used. The
NACA 0012-34 has a one-quarter leading-edge radius of the NACA
2. Paint Formulation and Calibration Methods 0012 and maximum thickness at 40% of the leading edge. Static
In this experiment, Pd(II) meso-tetra (pentafluorophenyl) porphine pressure taps were provided on the centerline on the upper surface to
(PdTFPP) and poly 1-trimethylsilyl propyne [poly(TMSP)] were measure the reference pressure for PSP data using a multichannel
used as the sensor molecule and binder, respectively. PSP composed pressure scanner (PSI Pressure Systems 9116), and a thermocouple
of PdTFPP and poly(TMSP) is known to have high-pressure was also installed inside the test model near the centerline to obtain
sensitivity under low-pressure conditions [19,20]; therefore, it is the temperature correction factor for PSP measurement. Five static
considered applicable to low-pressure experiments in the MWT. The pressure taps and one pressure tap were provided on the flat plate and
composition of the paint is PdTFPP (4.8 mg), poly(TMSP) (0.16 g), the NACA 0012-34, respectively. Each pressure tap is connected to a
and toluene (20 ml). The absorption peak is approximately 407 nm, multichannel pressure scanner through pressure tubes. These test
and the emission peak is 670 nm. models were set into the test section by adjusting the gap between the
Sample tests were conducted using chamber calibration [21]. The wing tip and the side wall to 0.15 mm. Rae and Pope [22] suggested
test gas is air, and the temperature is 20°C. Pressure sensitivity is that the gap should be lower than 0.5% of the span. Furthermore,
approximately 77%∕kPa in the pressure range below 1 kPa; this is Mueller and Burns [23] showed that the gap in the range from 0.1 to
Downloaded by KYUSHU UNIVERSITY OOA on July 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/1.J053468

sufficiently high to resolve a small change in pressure. It is noted 1.4 mm does not affect the results. Therefore, we determined the gap
that pressure sensitivity is dependent on temperature. Also, the effect size based on these previous studies.
of temperature on pressure sensitivity can be neglected because the
temperature change on the model is small during a test time of B. Experimental Condition
less than 1 min. Because of the effect of photodegradation caused We defined the experimental condition at a Mach number of 0.20
by continuous UV light irradiation, the luminescence intensity and a Reynolds number of 1.1 × 104 as a basic case. For evaluating
decreases by up to 5% for 30 min. The effect of photodegradation is pure Reynolds-number effects, the Mach number was fixed at ap-
considered negligible in MWT tests because the test time is usually proximately 0.20, and the Reynolds number was changed from
less than 1 min. At room temperature, the response time is about 1 s. 4.0 × 103 to 4.0 × 104 . By contrast, for evaluating pure Mach-
This is a required consideration for the response time in the MWT number effects, the Reynolds number was maintained at 1.1 × 104 ,
tests, especially under lower-pressure conditions, because the test and the Mach number was changed from 0.1 to 0.61. The Mach and
time becomes shorter with decreasing total pressure. Reynolds numbers were independently varied by controlling the total
To date, use of the PSP technique in the CO2 mode has not yet been pressure and flow velocity. A series of tests was conducted in the
established. Thus, the PSP measurement in this study is applied only range of the total pressures from 2 to 20 kPa. Air and CO2 were used
to air mode cases. as the test gases. The specific heat ratio was 1.4 in air and 1.3 in CO2 .
The driving and driven gas species are the same.
3. Setup for Pressure-Sensitive Paint Measurement
The setup for PSP measurement is shown in Fig. 3. The optical
equipment was placed outside the vacuum chamber, and mea- V. Estimation of Separation Bubble Characteristics
surements were performed through an optical window on the top of Gerakopulos [24] suggested an estimation approach of separation,
the vacuum chamber and a transparent ceiling of the test section made transition, and reattachment locations based on a pressure profile
of acrylic. Two UV–light-emitting diode (LED) arrays with 395 nm
wavelength were used as the excitation light source. A luminescent
image was captured by a thermoelectrically cooled 12 bit charge-
coupled device (CCD) camera (Hamamatsu, C4742-95-12ER) with
an optical bandpass filter (transmission wavelength 670 $ 20 nm).
An F1.4 camera lens (Nikon, 50 mm) was attached to the camera. The
resolution of the PSP image is approximately 0.14 mm. In this study,
calibration is made in the test section and using the model during the
evacuation process of the vacuum chamber. The change in model
temperature during the calibration is small and negligible. In this Fig. 4 Test models.
method, the calibration coefficients are obtained on each pixel of the
image of the model so that the nonuniformity in pressure sensitivity
can be corrected.

Fig. 5 Estimation of separation (S), transition (T), and reattachment (R)


Fig. 3 PSP measurement in the MWT. locations based on pressure profile.
ANYOJI ET AL. 1643

on the upper surface of an airfoil model. Figure 5 shows a typical VI. Results and Discussions
pressure distribution when a laminar separation bubble is formed. In A. Flat Plate
this method, the separation location is estimated as the intersection of 1. Basic Case
a linear fit to the nearly linear pressure recovery region and a linear fit
Figure 6 shows the lift and drag for the basic case at a Reynolds
to the nearly constant-pressure region. For estimating transition and
number of 1.1 × 104 and a Mach number of 0.20.
reattachment locations, a shape-preserving polynomial fit in the
The lift-curve slope is nearly linear up to α ! 6 deg and is very
regions of constant surface pressure and subsequent rapid surface close to "Cl ∕α# ! 2π for the two-dimensional potential flow lift-
pressure recovery is used. The transition location is estimated as the curve slope. Although the lift coefficient starts to level off at approx-
local maximum in the second derivative of the polynomial fit, imately α ! 7 deg, conventional stall characteristics, such as a drop
whereas the reattachment location is estimated as the location of the in lift coefficient as seen in a smooth airfoil at much higher Reynolds
local minimum in the second derivative of the polynomial fit. By numbers (Re > 105 ), are not observed at the angles of attack greater
applying the procedure to the mean pressure distributions obtained than α ! 8 deg. Furthermore, although a substantial decrease in
by the PSP measurement, locations of separation, transition, and lift from α ! 9 deg to α ! 11 deg is observed, the lift coefficient
reattachment are estimated. gradually increases with increasing angle of attack. Eventually, the
It is noted that the term “transition” is used in this paper based on maximum lift reaches 0.76 at α ! 15 deg, corresponding to the
Mueller’s interpretation; however, actually, we consider that a angle-of-attack limit in this experiment. This stall characteristic cor-
laminar-to-turbulent transition probably differs from a general tran- responds to the thin-airfoil stall.
sition of a boundary layer because flow phenomena occur in the Figure 7 shows PSP images on the upper surface and pressure
unstable separated shear layer. Thus, in this paper, “turbulence” is profiles on the centerline in the span direction for the basic case. The
Downloaded by KYUSHU UNIVERSITY OOA on July 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/1.J053468

defined as complex three-dimensional disturbed flows where the black dots plotted in Fig. 7 represent static pressure data obtained
two-dimensionality of the flow structure breaks. through static pressure taps as reference pressure. It is noted that,
in some cases, pressure distribution is slightly distorted at the model
free end, probably because of the existence of flow passing through
0.8 the narrow gap between the model tip and the wall (nominal 0.3 mm).
Because the low-pressure region decreases near the corners on the
upper surface, lift may become smaller due to the effect of the
0.6
three-dimensional flow as compared with that under a complete
two-dimensional flowfield. The effect of three-dimensionality on
0.4 aerodynamics should appear as a decrease of the lift-curve slope.
However, compared with the two-dimensional potential flow lift-
0.2 curve slope as shown in Fig. 6, a notable decrease and a non-
Cl and Cd

linearity of the lift-curve slope are not observed. Besides, the effect of
three-dimensionality is limited to a small region near the walls.
0
Therefore, the effects of the three-dimensional flow on the
aerodynamics can be considered negligible. This effect does not
–0.2 Cl affect our discussion on the flowfields based on the Cp distribution
Cd curves because two dimensionality of the flow is maintained in the
–0.4 center region.
2πα The low-pressure region appears around the leading edge and
–0.6 increases downstream with increasing angle of attack. Figure 8 shows
–5 0 5 10 15 locations of separation, transition, and reattachment with angles of
α [deg] attack for the basic case in accordance with Gerakopulos’s procedure.
Fig. 6 Aerodynamic characteristics of a flat plate in the basic case The low-pressure region seen in Fig. 7 corresponds to a laminar
(Re ! 1.1 × 104 , M ! 0.20). separation bubble. Flow separates at the leading edge for all angles of

-0.8 -0.8 -0.8 -0.8 -0.8

-0.6 -0.6 -0.6 -0.6 -0.6


Cp

Cp

-0.4
Cp

-0.4 -0.4
Cp

Cp

-0.4 -0.4

-0.2 -0.2 -0.2 -0.2 -0.2

0 0 0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x/c x/c x/c x/c x/c

a) α = 0 deg b) α = 3 deg c) α = 5 deg d) α = 7 deg e) α = 10 deg


Fig. 7 PSP images (top) and pressure distribution (bottom) of a flat plate for the basic case (black dots represent static pressure data obtained by a
pressure scanner).
1644 ANYOJI ET AL.

12 Separation from 4.0 × 103 to 4.0 × 104 , and the Mach number was maintained
Transition constant at 0.2. In all cases, lift curves are almost linear below
Reattachment α ! 5 deg to α ! 7 deg; however, the lift-curve slopes decrease
10 with decreasing Reynolds number. The lift-curve slopes (Cl ∕α) at
Re ! 4.0 × 104 and at Re ! 4.0 × 103 are 7.64 and 5.53∕rad,
8 respectively. This indicates that the lift-curve slope of a flat plate
decreases by 26.7% because of the difference in one order in the
α [deg]

6 Reynolds number. By contrast, the stall angles, which are defined as


the angle where the linearity of the lift starts collapsing, increase with
4 decreasing Reynolds number; that is, burst is less likely to occur at
lower Reynolds numbers. At the poststall angles of attack, lift
2
coefficients start to level off and maintain almost a constant value
up to α ! 15 deg, and the drag rapidly increases near the stall angles
in all cases.
0 We found that, although the drag coefficients are larger below
0 0.2 0.4 0.6 0.8 1.0 approximately α ! 3 deg for higher Reynolds numbers, the drag
x/c coefficients increase with decreasing Reynolds number above
Fig. 8 Variation of separation, transition, and reattachment locations approximately α ! 3 deg. It is noted that the absolute values of
with angle of attack for the basic case. measured drag are fairly small. The drag at zero angle of attack at
Downloaded by KYUSHU UNIVERSITY OOA on July 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/1.J053468

Re ! 1.1 × 104 , for instance, is approximately 0.05 N, which is


1∕120 of the load-cell range for the drag measurement. Because the
attack, whereas transition and reattachment locations move down- difference in drag at low angles of attack may be within the margin of
stream as angles of attack increase for α < 7 deg. This indicates that the error, we may avoid further discussion of this difference.
the separation bubble simply grows without the shift below α ! Figure 10 shows PSP images and pressure profiles on the flat plate
7 deg with increasing angles of attack. Note that the three dimen- for three angles of attack and three Reynolds numbers. The Mach
sional flow observed in Fig. 7, in particular at α ! 0 deg and α ! number was maintained constant at 0.20. At angles of attack less than
3 deg, is due to a corner vortex generated at the corner by the 3 deg, a low-pressure region around the leading edge corresponding to
separated flow. In contrast, the separated shear layer does not reattach leading-edge separation bubble is observed common for all Reynolds-
to the model surface over α ! 7 deg, where the lift coefficient starts number cases. However, the Cp profile shows different trends only at
to level off, as shown in Fig. 6. Thus, we define a stall angle of attack Re ! 4.9 × 103 compared with those at the other Reynolds numbers.
for a flat plate where the linearity of the lift starts collapsing because At α ! 0 deg for Re ! 4.9 × 103, shown in Fig. 10a, the pressure
of laminar-separation bubble burst. Flow visualization for Re ! slowly increases from the leading edge without a constant-pressure
10; 000 by Okamoto and Azuma [25] around a thin elliptical wing at region, unlike the other two cases, shown in Figs. 10d and 10g. The Cp
α ! 10 deg with AR ! 1, whose cross section is rectangular, indi- profile for Re ! 4.9 × 103 means that, although flow separates at the
cates unsteady vortices shedding from the leading edge without clear leading edge, and the separated shear layer reattaches to the surface at
reattachment. Although two- and three-dimensional wing shapes approximately x∕c ! 0.20, a laminar–turbulent transition does not
differ between their experiments and ours, we consider that the occur, which is a profound difference. Sasaki and Kiya [27] visualized
flowfields around the cross section at the span centerline are re- the vortex structure on the flat plate at various low Reynolds numbers.
markably similar. In addition, Garmann and Visbal [26] showed that a Their visualization result at Re ! 5.6 × 103 , which is close to our
negative pressure was maintained on the flat plate wing even after experimental condition (Re ! 4.9 × 103 ), shows that the separated
breakdown of a coherent leading-edge vortex (LEV), and a rapid drop shear layer remains laminar up to the reattachment location. Further-
in lift coefficient was not found before and after breakdown of the more, surface pressure measurements for the flow over a backward-
LEV. Hence, we deduce that a low-pressure region on the flat plate as facing step were conducted by Tani et al. [28]. They investigated that
shown in Figs. 7d and 7e is created by the unsteady leading-edge the separated flow at the corner does not undertake a transition and
separation vortices and prevents a sharp decrease in the lift coefficient reattaches in a laminar state when the step height is very small, and
even at the poststall angle of attack due to the vortex lift. then the pressure on the surface slowly recovers without a steep
pressure recovery due to a laminar-to-turbulent transition. The pres-
2. Reynolds-Number Effect sure profile shown in Fig. 10a reveals the feature very similar to the
Figure 9 shows the effects of the Reynolds number on lift and drag moderate pressure recovery in the step flow observed by Tani et al. The
characteristics of the flat plate. The Reynolds number was varied results of these past studies agree with our results and explanation.

0.8
0.8
0.6

0.4
0.4
0.2
Re = 4.0 × 104
Cl

0
Cl

4
0 Re = 4.0 × 104
Re = 2.0 × 10 4
–0.2 4 Re = 2.0 × 10
Re = 1.1 × 10 4
3 Re = 1.1 × 10
–0.4 Re = 6.0 × 10 –0.4
3
3 Re = 6.0 × 10
–0.6 Re = 4.0 × 10 3
Re = 4.0 × 10

–0.8 –0.8
–5 0 5 10 15 0 0.05 0.10 0.15 0.20 0.25 0.30 0.35
α [deg] Cd

a) Lift curve b) Drag polar


Fig. 9 Effects of the Reynolds number on aerodynamic characteristics of a flat plate.
ANYOJI ET AL. 1645

-1.2 -1.2 -1.2


-1 -1 -1
-0.8 -0.8 -0.8
-0.6 -0.6
Cp

Cp
-0.6

Cp
-0.4 -0.4 -0.4
-0.2 -0.2 -0.2
0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x/c x/c x/c
3 3 3
a) Re = 4.9×10 , α = 0 deg b) Re = 4.9×10 , α = 3 deg c) Re = 4.9×10 , α = 4 deg

-1.2 -1.2 -1.2


-1 -1 -1

-0.8 -0.8 -0.8

-0.6 -0.6 -0.6

Cp
Cp

Cp
-0.4 -0.4 -0.4

-0.2 -0.2 -0.2


Downloaded by KYUSHU UNIVERSITY OOA on July 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/1.J053468

0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x/c x/c x/c
d) Re = 1.1×104, α = 0 deg e) Re = 1.1×104, α = 3 deg f) Re = 1.1×104, α = 5 deg
-1.2 -1.2 -1.2
-1 -1 -1
-0.8 -0.8 -0.8
Cp

-0.6 -0.6 -0.6


Cp

Cp
-0.4 -0.4 -0.4
-0.2 -0.2 -0.2
0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x/c x/c x/c
g) Re = 4.1×104, α = 0 deg h) Re = 4.1×104, α = 3 deg i) Re = 4.1×104, α = 5 deg
Fig. 10 Pressure distributions on the upper surface of a flat plate at Re ! 4.9 × 10 (top), Re ! 1.1 × 104 (center), and Re ! 4.1 × 104 (bottom).
3

In correlation with these results, the PSP data suggest that a critical poststall angles of attack only for M ! 0.61 and M ! 0.47 slowly
region, where the vortex structure around the flat plate at α ! 0 deg increase. The largest difference in lift, at α ! 15 deg between each
changes, exists between Re ! 4.9 × 103 and Re ! 1.1 × 104 . Mach number condition, is approximately 9.8%. Also, almost no
variation is seen in the polar curves, as shown in Fig. 11b. Hence, the
3. Mach-Number Effect effect of the Mach number on aerodynamic characteristics of a flat
Figure 11 shows the effects of the Mach number on the lift and drag plate in Mach number up to 0.61 is almost negligible, whereas we
characteristics of the flat plate model for a constant Reynolds number. may draw attention to the lift coefficient at high angles of attack above
The Mach number was varied from 0.09 to 0.61, whereas the α ! 15 deg because there is a possibility that the Mach number
Reynolds number was maintained constant at 1.1 × 104 . Lift curves increases the difference in lift.
are nearly unchanged even when the Mach number changes up to Figure 12 shows PSP images and pressure profiles on the flat plate
0.61 compared with that in the basic case (M ! 0.20). The lift-curve for three angles of attack and three Mach numbers. The Reynolds
slope slightly increases at M ! 0.61, and the lift coefficients at the number was maintained constant at 1.1 × 104 . In all cases, a leading-

1.0 1.0
0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2
Cl

Cl

0 0

–0.2 M = 0.61 –0.2 M = 0.61


M = 0.47 M = 0.47
–0.4 M = 0.33 –0.4 M = 0.33
M = 0.20 M = 0.20
–0.6 M = 0.09 –0.6 M = 0.09
–5 0 5 10 15 0 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40
α [deg] Cd
a) Lift curve b) Drag polar
Fig. 11 Effect of Mach number on aerodynamic characteristics of a flat plate.
1646 ANYOJI ET AL.

edge separation bubble with a low-pressure region and similar Cp affected by the difference in specific heat ratio in the Reynolds
profiles are observed. However, focusing on transition and re- number region above Re ! 1.1 × 104 .
attachment locations defined on the basis of Cp profiles, these loca- Figure 14 shows the effect of the Mach number on the lift and drag
tions move downstream as the Mach number increases. That is, the characteristics of the flat plate in different specific heat ratios. The
effect of compressibility stabilizes the separated shear layer and then Reynolds number was constant at 1.1 × 104 , and the Mach number
delays both laminar–turbulent transition and reattachment. Conse- was varied from 0.20 to 0.61. Although a slight difference exists in the
quently, a separation bubble is slightly extended. However, this flow lift coefficient at the poststall angle of attack between the air and CO2
change has a very small effect on aerodynamic characteristics modes, almost no variation exists in the lift-curve slope and drag
because of a negligible change in Cp profiles. characteristics. Comparing the lift coefficients after stall between the
air and CO2 modes at the same Mach number, the lift in the CO2 mode
4. Effect of Specific Heat Ratio is slightly lower than that in the air mode, in common with the previous
Figure 13 shows the Reynolds number effect on lift and drag case shown in Fig. 13. However, the difference in the lift coefficient
characteristics of the flat plate in different specific heat ratios (air and between the air and CO2 modes is 4.3%, maximum. Hence, the effect
CO2 ). The Reynolds number was varied from 4.0 × 103 to 4.0 × 104 , of the Mach number on specific heat ratio is almost negligible.
and the Mach number was maintained constant at 0.20.
In the Reynolds-number range above 1.1 × 104 , the lift slope, stall B. NACA 0012-34
angle of attack, and drag polar change little in the air and CO2 modes. 1. Basic Case
However, a difference exists in the lift coefficient at the poststall angle Figure 15 shows the aerodynamic characteristics of the NACA
of attack. These differences in the lift coefficient between the air and 0012-34 airfoil for the basic case (Re ! 1.1 × 104 , M ! 0.20). A
Downloaded by KYUSHU UNIVERSITY OOA on July 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/1.J053468

CO2 modes are not common among different Reynolds numbers. For strong nonlinearity of the lift curve and a rapid decrease of the lift
the smallest Reynolds number, the lift slope and maximum lift after stall are observed, unlike the characteristics of the lift curve for
coefficient in the CO2 mode were larger than those in the air mode. the flat plate as described previously. Although the lift-curve slope in
The difference may be caused by the higher Reynolds number in the the range from α ! 0 deg to α ! 5 deg is no more than Cl ∕α !
CO2 mode because the lift slope has a tendency to become large with 2.04∕rad, the lift coefficient drastically increases from α ! 7 deg to
increasing Reynolds number. This Reynolds number difference α ! 8 deg. The lift coefficient at α ! 8 deg achieves approximately
between the air and CO2 modes is 22.5%, which is high compared 1.7 times the lift coefficient at α ! 7 deg by only 1 deg increase of
with the difference in the other Reynolds number conditions between the angle of attack. In addition, a drag bucket is observed from α !
air and CO2 modes. Therefore, a comparison of the effects of the 7 deg to α ! 9 deg. The lift coefficient moderately increases above
specific heat ratio cannot be discussed accurately only for Reynolds α ! 8 deg and then reaches a peak at α ! 11 deg, corresponding
numbers below Re ! 1.1 × 104 ; however, at the very least, it can be to the stall angle of attack. It keeps on decreasing after the stall up
said that the aerodynamic performance of the flat plate is little to α ! 15 deg.

-1 -1 -1

-0.8 -0.8 -0.8

-0.6 -0.6 -0.6


Cp

Cp
Cp

-0.4 -0.4 -0.4

-0.2 -0.2 -0.2

0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x/c x/c x/c
a) M = 0.21, α = 0 deg b) M = 0.21, α = 3 deg c) M = 0.21, α = 5 deg

-1 -1 -1

-0.8 -0.8 -0.8

-0.6 -0.6 -0.6


Cp
Cp

Cp

-0.4 -0.4 -0.4

-0.2 -0.2 -0.2

0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x/c x/c x/c
d) M = 0.47, α = 0 deg e) M = 0.47, α = 3 deg f) M = 0.47, α = 5 deg

-1 -1 -1
-0.8 -0.8 -0.8
-0.6 -0.6 -0.6
Cp

Cp

Cp

-0.4 -0.4 -0.4


-0.2 -0.2 -0.2
0 0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x/c x/c x/c
g) M = 0.60, α = 0 deg h) M = 0.60, α = 3 deg i) M = 0.60, α = 5 deg
Fig. 12 Pressure distributions on the upper surface of a flat plate for three angles of attack and three Mach numbers.
ANYOJI ET AL. 1647

1
0.8

0.6
0.5
0.4 Air mode Air mode
CO2 mode CO2 mode
0.2
4
Re = 4.0 × 104 0 Re = 4.0 × 10

Cl
Cl
0 4 4
Re = 1.1 × 10 Re = 1.1 × 10
3 3
–0.2 Re = 4.0 × 10 Re = 4.0 × 10
4 4
Re = 3.8 × 10 –0.5 Re = 3.8 × 10
–0.4
4 4
Re = 1.1 × 10 Re = 1.1 × 10
–0.6 3 Re = 4.9 × 10
3
Re = 4.9 × 10
–0.8 –1
–5 0 5 10 15 0 0.05 0.10 0.15 0.20 0.25 0.30 0.35
α [deg] Cd

a) Lift curve b) Drag polar


Fig. 13 Effects of specific heat ratio on aerodynamic characteristics of a flat plate for constant Mach number (M ! 0.2).
Downloaded by KYUSHU UNIVERSITY OOA on July 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/1.J053468

1 1

0.5
0.5

Air mode 0 Air mode


Cl

Cl

0 CO2 mode CO2 mode

M = 0.61 M = 0.61
M = 0.49 M = 0.49
M = 0.20 –0.5 M = 0.20
–0.5 M = 0.60 M = 0.60
M = 0.47 M = 0.47
M = 0.20 M = 0.20
–1
–5 0 5 10 15 0 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40
α [deg] Cd
a) Lift curve b) Drag polar
Fig. 14 Effects of specific heat ratio on aerodynamic characteristics of a flat plate for a constant Reynolds number (Re ! 1.1 × 104 ).

Figure 16 shows PSP images and Cp profiles of the NACA 0012- seem to reattach to the surface at α ! 7 deg; however, it is difficult to
34 airfoil for the basic case. Figure 17 shows a variation of separation, clearly estimate whether the separated shear layer reattaches based
transition, and reattachment locations based on Cp analysis. Details only on the feature of Cp profiles. At α ! 8 deg, the separation
about the flowfield are described later in this paper. A trailing-edge location moves upstream and then passes over the high point on the
separation occurs at x∕c ! 0.95 for α ! 0 deg. The separation point airfoil (x∕c ! 0.40). Furthermore, a laminar separation bubble forms
moves upstream to x∕c ! 0.53 for α ! 6 deg, and a suction peak in the vicinity of the center of the chord length. The formation of the
also forms at the leading edge. At α ! 7 deg, a laminar-to-turbulent separation bubble enhances the negative pressure on the upper
transition occurs at x∕c ! 0.72. The separated shear layer does not surface and consequently causes a nonlinear lift increase, as shown in

0.8 0.8

0.6 0.6

0.4 0.4
Cl
Cl

0.2 0.2

0 0

–0.2 –0.2
0 0.05 0.10 0.15 0.20 0.25 0.30 0.35
–5 0 5 10 15 20
Cd
α [deg]
a) Lift curve b) Drag polar
Fig. 15 Aerodynamic characteristics of the NACA 0012-34 airfoil in the basic case (Re ! 1.1 × 104 , M ! 0.20).
1648 ANYOJI ET AL.

-2 -2 -2 -2 -2

-1.5 -1.5 -1.5 -1.5 -1.5

-1 -1 -1 -1
Cp

Cp

Cp

Cp
-1

Cp
-0.5 -0.5 -0.5 -0.5
-0.5
Downloaded by KYUSHU UNIVERSITY OOA on July 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/1.J053468

0 0 0 0
0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x/c x/c x/c x/c x/c
a) α = 0 deg b) α = 6 deg c) α = 7 deg d) α = 8 deg e) α = 10 deg
Fig. 16 Pressure distribution of the NACA 0012-34 for the basic case (black dot represents static pressure data) obtained by a pressure scanner.

Fig. 15. The separation bubble moves a bit upstream up to α ! strongly sensitive to the Reynolds number change. The most
10 deg as the angle of attack increases. At α ! 14 deg, corre- pronounced common feature is a high nonlinearity of the lift curves.
sponding to the poststall angle of attack, reattachment is not observed The kink point, meaning the angle of attack where the lift coefficient
in the Cp profile; however, the transition is observed at starts to sharply increase, delays with decreasing Reynolds number.
around x∕c ! 0.30. Each kink point corresponds to α ! 3 deg, α ! 7 deg, and α !
Contrary to the flat plate, the effect of three-dimensionality on Cp 11 deg for Re ! 3.9 × 104, Re ! 1.1 × 104 , and Re ! 4.0 × 103 ,
distribution for the NACA 0012-34 airfoil shown in Fig. 16 is not respectively. It is noted that the lift curves overlap well in all the
found at α ! 0 deg because a large area of the airfoil surface is Reynolds numbers before a nonlinear lift increase. The lift coefficient
covered by the attached flow. In the angle-of-attack range from α ! at the angles of attack after the steep lift enhancement gradually
0 deg to α ! 5 deg, where the flow attaches to the surface in large increases with the angle of attack and then reaches a peak point
area of the wing, we consider that the effect of three-dimensionality on corresponding to the stall condition. However, differences in the stall
the aerodynamics is considered to be small due to the two dimen- angle and maximum lift coefficient Cl max are found. The stall angle
sionality of the flow. However, above α ! 6 deg, where the sep- for 1.1 × 104 is α ! 11 deg, whereas the peak of lift coefficient
aration point moves upstream, the two-dimensional nonuniformity of seems to be approximately at α ! 14.5 deg for Re ! 4.0 × 103. In
Cp distribution appears downstream of the separation point. As with addition, Cl max gradually decreases as the Reynolds number de-
the discussion on the flat plate, the effect on aerodynamics is negligible creases. A 13.3% decrease of Cl max from Re ! 3.9 × 104 to Re !
even if the three-dimensionality appears because the two-dimensional 4.0 × 103 is observed. In drag polars shown in Fig. 18b, a drag bucket
uniformity is limited to a small region near the walls. is found both at Re ! 3.9 × 104 and Re ! 1.1 × 104 , as seen in the
basic case. The drag increases with decreasing Reynolds number,
2. Reynolds-Number Effect
indicating that the effect of viscosity becomes pronounced. The drag
at Re ! 3.9 × 104 changes little even when the angle of attack
Figure 18 shows the effects of the Reynolds number on the lift and increases up to α ! 8 deg. By contrast, the drag gradually increases
drag characteristics of the NACA 0012-34 airfoil model. In contrast as the angle of attack increases at Re ! 4.0 × 103 .
to the results of the flat plate (Fig. 9), the lift curves and drag polars are These unusual aerodynamic variations with nonlinearity as
displayed in Fig. 18 are associated with the behavior of a separation
20 bubble. Figure 19 shows the pressure distribution at Re ! 4.9 × 103
and 4.1 × 104 . Figure 20 presents the variation of separation, tran-
sition, and reattachment locations based on the analysis of Cp profiles
applying Gerakopulos’s methodology, and Fig. 21 illustrates
15 schematic images of flowfields on the airfoils for each Reynolds
number.
Although the laminar separation without reattachment occurs
near the trailing edge at α ! 0 deg for Re ! 4.9 × 103, similar
α [deg]

10 trends of Cp distributions at Re ! 4.9 × 103 are observed at from


α ! 9 deg to α ! 11 deg, where the pressure distribution has a
suction peak at the leading edge and a constant-pressure region. This
indicates that flow separation occurs without a laminar-to-turbulent
5 transition. The flow remains separated up to α ! 11 deg, and the
Separation separation location gradually moves upstream as the angle of attack
Transition increases. This separated flow leads to a lower lift-curve slope than
Reattachment those for the other higher Reynolds numbers. However, at α ! 12
0
0 0.2 0.6 0.4
0.8 1.0 and 14 deg, the pressure recovery is found at around x∕c ! 0.80,
x/c indicating that at least the transition occurs. Although it is difficult to
Fig. 17 Variation of separation, transition, and reattachment locations conclusively determine whether or not the flow reattaches based only
with the angle of attack for the basic case. on the shape of Cp distribution, the reattachment location at α ! 12
ANYOJI ET AL. 1649

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

Cl
Cl
0 0

–0.2 –0.2
Re = 3.9 × 104 Re = 3.9 × 104
–0.4 Re = 1.1 × 10
4
–0.4 Re = 1.1 × 10
4
3
Re = 4.0 × 10 Re = 4.0 × 10
3

–0.6 2π –0.6
0 0.05 0.10 0.15 0.20 0.25 0.30 0.35
–5 0 5 10 15 20
Cd
α [deg]
a) Lift curve b) Drag polar
Fig. 18 Effects of the Reynolds number on aerodynamic characteristics of the NACA 0012-34 airfoil for a constant Mach number (M ! 0.19 to 0.22).
Downloaded by KYUSHU UNIVERSITY OOA on July 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/1.J053468

and 14 deg is mathematically worked out by applying Gerakopulos’s separation bubble forms at α ! 3 deg and causes a rapid lift increase.
methodology. Besides, Fig. 18 reveals a nonlinear lift increase In this angle of attack, the separation point slightly moves upstream
between α ! 11 deg and α ! 12 deg, suggesting a lift enhance- from x∕c ! 0.76 at α ! 0 deg, and then the separated shear layer
ment due to a formation of a separation bubble. The separation bubble reattaches near the trailing edge. The separation bubble continues to
covers a wide range of the wing from α ! 12 deg to α ! 14 deg. At slightly move upstream up to α ! 5 deg. By contrast, the location
α ! 16 deg, the pressure distribution denotes the same tendency at of the separation bubble drastically changes from α ! 5 deg to
α ! 11 deg, indicating that only the flow separation occurs without α ! 6 deg. The separation bubble passes over the high point of the
transition and reattachment, and the burst of the separation bubble wing at α ! 6 deg, and a comparatively short separation bubble
occurs. A small drop in lift is also observe between α ! 15 deg and forms near the leading edge. However, the shift of the separation
α ! 16 deg in Fig. 18. It seems that lift coefficient reaches its peak at bubble has little effect on the nonlinearity of the lift curve. Fur-
approximately α ! 14.5 deg. Hence, we concluded that the burst of thermore, the difference in the drag coefficient between α ! 5 deg
the separation bubble occurs at α ! 16 deg, and the stall angle is and α ! 6 deg is very small. As shown in Fig. 19b, the negative
α ! 14.5 deg. Alam et al. [29] and Zhou et al. [30] inves- pressure around the leading edge (x∕c ! 0–0.15) increases due to the
tigated aerodynamics and the wake of the NACA 0012 airfoil at a forward movement of the separation bubble. However, this negative
similar Reynolds number. Their results showed that flow remained pressure region of the separation bubble has little contribution to drag
laminar after separation without reattachment at Re ! 5.3 × 103 . In because the chordwise location of x∕c ! 0.17 corresponds to the
addition, the lift coefficient has no peak value, indicating the location where the slope of the tangent line to the airfoil profile is zero
conventional stall is absent. These results have a different tendency at α ! 6 deg; that is, the drag component at around x∕c ! 0.17 is
from our results. Many papers [4,6,7,31] have reported that aero- negligible. Moreover, the magnitude of Cp values on the suction side
dynamic performance is highly dependent on the airfoil shape and does not drastically change between α ! 5 deg and α ! 6 deg.
Reynolds number at such low Reynolds numbers. The NACA 0012 Although Cp distribution on the pressure side is also required for
airfoil has a comparatively large leading-edge radius, and thus the discussing about lift and drag in detail, we deduce that this is the
laminar boundary layer is more likely to separate. In contrast, reason why the differences in lift and drag coefficient between α !
the NACA 0012-34 airfoil has a smaller leading-edge radius, and the 5 deg and α ! 6 deg are small even if the separation bubble dras-
flow is less likely to separate. We consider that the difference between tically moves upstream. At α ! 12 deg, the flow separates at the
our results and those by Alam et al. [29] and Zhou et al. [30] is due to leading edge without reattachment.
the difference in the airfoil shape. Consequently, the behaviors of the separation bubble especially in
The flowfields at Re ! 1.1 × 104 (basic case) is as described in formation and burst of the separation bubble are considerably
Sec. VI.B.1. At Re ! 4.1 × 104 , a laminar separation without re- sensitive to the Reynolds number change and consequently lead to
attachment occurs at x∕c ! 0.76 for α ! 0 deg. However, the the distinguishing nonlinear lift curves at each Reynolds number.

-2.5 -2.5
= 0 deg = 0 deg
-2 = 5 deg -2 = 3 deg
= 9 deg = 5 deg
= 11 deg = 6 deg
-1.5 = 12 deg -1.5 = 8 deg
= 14 deg = 12 deg
= 16 deg
-1 -1
Cp

Cp

-0.5 -0.5

0 0

0.5 0.5
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x/c x/c
a) Re = 4.9 103 b) Re = 4.1 104
Fig. 19 Pressure distribution of the NACA 0012-34 at Re ! 4.9 × 103 and 4.1 × 104 .
1650 ANYOJI ET AL.

20 20 20

15 15 15

[deg]
[deg]

[deg]
10 10 10

5 5 5
Separation Separation Separation
Transition Transition Transition
Reattachment Reattachment Reattachment
0 0 0
0 0.2 0.4 0.6 0.8 1.0 0 0.2 0.4 0.6 0.8 1.0 0 0.2 0.4 0.6 0.8 1.0
x/c x/c x/c
a) Re = 4.9 103 b) Re = 1.1 104 c) Re = 4.1 104
Fig. 20 Variation of separation, transition, and reattachment locations with the angle of attack for different Reynolds numbers.

3. Mach-Number Effect Figure 23 shows the pressure distribution at M ! 0.48 and


Downloaded by KYUSHU UNIVERSITY OOA on July 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/1.J053468

Figure 22 shows the effect of the Mach number on the lift and drag M ! 0.61. Figure 24 presents a variation of separation, transition,
characteristics of the NACA 0012-34 airfoil model for a constant and reattachment locations. The Mach number was varied from 0.21
Reynolds number. Below α ! 6 deg, the lift curves and drag polars to 0.61, and the Reynolds number was maintained constant at
change little with increasing angles of attack. Hence, the Mach- 1.1 × 104 . The flowfield images, based on the information in Fig. 24,
number change does not affect aerodynamic performance in the are shown schematically in Fig. 25. In all cases, a laminar separation
low-angle-of-attack range. However, there is a noticeable change of without reattachment occurs near the trailing edge at the angles
the lift coefficient in the range of α ! 7 deg to α ! 14 deg. The lift of attack below α ! 6 deg. Furthermore, the separation location
coefficients for M ! 0.21 and M ! 0.48 keep increasing with in- moves upstream as the angle of attack increases. The common trend
creasing angles of attack up to α ! 10 deg or α ! 11 deg, whereas that is a trailing-edge separation leads the overlap of lift coefficients
the lift coefficient at M ! 0.61 starts to level off at approximately below α ! 6 deg. At α ! 7 deg, the separation bubble forms only
α ! 10 deg. In this angle-of-attack range, the lift coefficients de- for M ! 0.21, whereas the flow remains separated for M !
crease as the Mach number increases. Compared with the lift coef- 0.48 and M ! 0.61. The separation bubble forms at α ! 8 deg for
ficients at α ! 10 deg for each Mach number, a 25.4% decrease of M ! 0.48. This 1 deg difference between M ! 0.21 and M ! 0.48
the lift coefficient from M ! 0.21 to M ! 0.61 is found. Each lift affects the delay of the kink point shown in Figs. 24 and 25. It is noted
coefficient overlaps at α ! 15 deg and maintains almost the same that, although we predicted a reattached flow at α ! 8 deg for
level above α ! 15 deg. M ! 0.48 due to the formation of the separation bubble based on the
The trend that the lift coefficients from α ! 7 deg to α ! 14 deg Cp profile, a steep lift increase at α ! 8 deg is not yet seen in the lift
decrease with increasing Mach number is different from those curve shown in Fig. 22. We deduce that the angle of attack of around
predicted by the Prandtl–Glauert rule. According to the Prandtl– 8 deg for M ! 0.48 is a very sensitive region regarding whether the
Glauert rule, lift coefficients increase due to the compressibility. flow reattaches, and a slight error of the angle of attack affects the
Similar results were observed by Suwa et al. [32], who investigated difference between the interpretation of the flowfield and the lift
Mach-number effects on the aerodynamic characteristics of a triangle curve. Meanwhile, the flow does not reattach up to α ! 13 deg for
airfoil at Re ! 10; 000. Their results show that the lift curves overlap M ! 0.61, and a separation bubble does not form. The nonexistence
each other at low angles of attack, whereas lift coefficients from α ! of a laminar separation bubble leads to lower lift coefficients for
5 deg to α ! 12 at M ! 0.70 are smaller than those at M ! 0.50 and M ! 0.61 at high angles of attack greater than α ! 7 deg.
M ! 0.15. Hence, our results indicate that the Prandtl–Glauert rule is As shown in Figs. 24 and 25, separation, transition, and re-
not applicable to such low-Reynolds-number conditions. attachment locations move downstream as the Mach number

Fig. 21 Flow model of the NACA 0012-34 airfoil for different Reynolds numbers.
ANYOJI ET AL. 1651

0.8 0.8

0.6 0.6

0.4 0.4
Cl

Cl
0.2 0.2

0 0
M = 0.61 M = 0.61
M = 0.48 M = 0.48
M = 0.20 M = 0.20
0.2 0.2
5 0 5 10 15 20 0
0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40
[deg] Cd
a) Lift curve b) Drag polar
Fig. 22 Effects of the Mach number on aerodynamic characteristics of the NACA 0012-34 for a constant Reynolds number (Re ! 1.1 × 104 ).
Downloaded by KYUSHU UNIVERSITY OOA on July 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/1.J053468

increases, particularly at high angles of attack above α ! 7 deg, that for the NACA 0012 at much smaller Mach number and
where the difference in lift curves is observed for each Mach number. Re ! 5.3 × 103 . For the NACA 0012-34, although the transition
This indicates that the compressibility has the effect of stabilizing the occurs, the reattachment location passes over the trailing edge at
separated shear layer and delaying separation and transition to M ! 0.61; that is, the separated flow is observed. Consequently, the
turbulence. Therefore, we consider that the reattachment location flowfield around the NACA 0012-34 is remarkably similar to that for
passes over the trailing edge for M ! 0.61 because of the com- the NACA 0012 at a much smaller Mach number and Re ! 5.3 × 103
pressibility effect, although the transition occurs. As a result, this in spite of a large difference in Mach number. It should be noted here
prevents the formation of the separation bubble. that lift keeps increasing above α ! 10 deg in common with each
Compared with results by Alam et al. [29] and Zhou et al. [30], other. As Alam et al. [29] and Zhou et al. [30] visualized, a coherent
there is a similarity of the lift curve at M ! 0.61 shown in Fig. 22a to vortex exists on the airfoil. This large separated vortex generates a

-2 -2
= 0 deg = 0 deg
= 6 deg = 6 deg
-1.5 = 7 deg -1.5 = 7 deg
= 8 deg = 8 deg
= 10 deg = 10 deg
-1 = 11 deg -1 = 13 deg
Cp

Cp

-0.5 -0.5

0 0

0.5 0.5
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x/c x/c
a) M = 0.48 b) M = 0.61
Fig. 23 Pressure distribution of the NACA 0012-34 at a) M ! 0.48, and b) M ! 0.61.

15 15 15

10 10 10
[deg]

[deg]

[deg]

5 5 5

Separation Separation Separation


Transition Transition Transition
Reattachment Reattachment Reattachment
0 0 0
0 0.2 0.4 0.6 0.8 1.0 0 0.2 0.4 0.6 0.8 1.0 0 0.2 0.4 0.6 0.8 1.0
x/c x/c x/c
a) M = 0.21 b) M = 0.48 c) M = 0.61
Fig. 24 Variation of separation, transition, and reattachment locations with the angle of attack for different Mach numbers.
1652 ANYOJI ET AL.
Downloaded by KYUSHU UNIVERSITY OOA on July 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/1.J053468

Fig. 25 Flow model of the NACA 0012-34 for different Mach numbers.

low-pressure region on the suction side and enhances lift as vortex difference between the air and CO2 modes is 22.5% in the smallest
lift. We deduce that this is because lift keeps increasing at high angles Reynolds-number case. Therefore, a precise discussion about the
of attack. effect of the difference of the Reynolds number is precluded only for
In contrast to the results for the flat plate as described in the smallest Reynolds-number case.
Sec. VI.A.3, the effect of compressibility on the aerodynamics is The nonlinear lift curves, including the kink points, and the drag
clearly observed at high angles of attack. As for the flat plate, the polars change little between the air and CO2 modes. However, a
separation point is fixed at the leading edge, and compressibility difference in the lift coefficient around the stall angle of attack is
affects the locations of the transition and the reattachment. Moreover, found. As seen in the results of the flat plate described in Sec. VI.A.4,
there is little difference in the magnitude of the Cp values between the lift coefficients near the stall angle of attack (α ! 11 deg) for
M ! 0.20 and 0.60 at the same angle of attack. Hence, compress- Re ! 3.9 × 104 and Re ! 1.1 × 104 in the air mode are larger than
ibility has little effect on the aerodynamics. By contrast, compress- those in the CO2 mode. However, only for Re ! 3.9 × 104, the trend
ibility for the NACA 0012-34 affects not only transition and was reversed, and the lift coefficient in the air mode becomes lower
reattachment but separation, indicating that the location and the than that in the CO2 mode. As a common trend in each Reynolds
length of the separation bubble moves on the suction side. In number, the stall angle of attack in the CO2 mode is slightly smaller
particular at high angles of attack, the flow does not reattach to the than that in the air mode.
surface at larger Mach number. Therefore, compressibility for the As with the conclusion obtained in the flat-plate case, the differ-
NACA 0012-34 affects the aerodynamics at high angles of attack ence in specific heat ratio has little effect on the aerodynamic charac-
because compressibility especially has an effect on reattachment. teristics when the Reynolds number is varied; however, it may affect
the flowfield after stall.
4. Effect of Specific Heat Ratio Figure 27 shows the effects of specific heat ratio on the lift and drag
Figure 26 shows the effect of specific heat ratio on the lift and drag characteristics of the NACA 0012-34 model for a constant Reynolds
characteristics of the NACA 0012-34 model for a constant Mach number. The Mach number was varied from 0.20 to 0.62, and the
number. The Reynolds number was varied from 4.0 × 103 to Reynolds number was maintained constant at 1.1 × 104 .
3.9 × 104 , and the Mach number was maintained constant at 0.20. As The lift–drag characteristics are well accorded between the air and
is the case with the results of the flat plate, the Reynolds-number CO2 modes except around the stall angle of attack. The maximum lift

0.8 0.8
0.6 0.6
0.4 0.4
Air mode Air mode
0.2 CO2 mode 0.2 CO2 mode
Cl

Cl

0 Re = 3.9 104 0 Re = 3.9 104


4 4
Re = 1.1 10 Re = 1.1 10
0.2 Re = 4.0 10
3 0.2 Re = 4.0 10
3

4 4
Re = 3.8 10 Re = 3.8 10
0.4 4
0.4 4
Re = 1.1 10 Re = 1.1 10
3 3
0.6 Re = 4.9 10 0.6 Re = 4.9 10

10 5 0 5 10 15 20 25 0 0.05 0.10 0.15 0.20 0.25 0.30 0.35


[deg] Cd
a) Lift curve b) Drag polar
Fig. 26 Effects of specific heat ratio on aerodynamic characteristics of the NACA 0012-34 for a constant Mach number (M ! 0.20).
ANYOJI ET AL. 1653

0.8 0.8

0.6 0.6

0.4 0.4

Cl
Cl

Air mode Air mode


0.2 CO2 mode 0.2 CO2 mode
M = 0.61 M = 0.61
M = 0.48 M = 0.48
0 M = 0.20 0 M = 0.20
M = 0.62 M = 0.62
M = 0.49 M = 0.49
M = 0.20 0.2 M = 0.20
0.2
10 5 0 5 10 15 20 25 0 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40
[deg] Cd
a) Lift curve b) Drag polar
Downloaded by KYUSHU UNIVERSITY OOA on July 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/1.J053468

Fig. 27 Effects of specific heat ratio on aerodynamic characteristics of the NACA 0012-34 for a constant Reynolds number (Re ! 1.1 × 104 ).

at the stall angle of attack in the CO2 case slightly deteriorates for compressibility on the flowfields around the flat plate does not
Mach numbers below 0.49. However, both lift and drag char- affect aerodynamic performance.
acteristics agree rather well in a wide range of the angles of attack For the NACA 0012-34 airfoil, the effects of Reynolds and Mach
between the air and CO2 modes for M ! 0.61, although a small numbers on aerodynamic characteristics become more prominent
difference is observed at α ! 12 deg and α ! 13 deg. Con- compared to the flat plate. The Reynolds-number effect leads to
sequently, effects of specific heat ratio on aerodynamic character- nonlinear aerodynamic variation such as a steep lift increase caused
istics are not large and become smaller as Mach number increases. by the behavior of the laminar separation bubble. The lift coefficient
Anyoji et al. [33] numerically investigated the flowfield and the rapidly increases at a certain angle of attack where the separated shear
aerodynamic performance of the NACA 0012 at Re ! 23; 000 and layer reattaches on the model surface, and the laminar separation
M ! 0.20. Their results showed that the spanwise vortices break up bubble forms. By contrast, the Mach-number effect decreases the
near the transition point at α ! 9 deg. At the downstream of the lift coefficient at high angles of attack at greater than 7 deg, whereas
transition point, the flow has strong three-dimensionality, and the lift curves for each Mach number change little with increasing angles
mixing of the separated shear layer is enhanced. Although the Mach of attack below 6 deg. This trend is ruled out by Prandtl–Glauert rule,
number in the CO2 mode is consistent with that in the air mode, the indicating that the Prandtl–Glauert rule is not applicable to such
flow velocity in the CO2 mode is lower than that in the air mode due to low-Reynolds-number conditions. In analogy with the flat plate,
the difference in sound velocity, and then the Reynolds stress in the the compressibility effect, which delays separation, transition, and
CO2 mode becomes smaller compared with the air mode. Hence, we reattachment, was observed. Effects of specific heat ratio on aero-
consider that the difference in momentum exchange by Reynolds dynamic characteristics are not large, in particular at angles of attack
stress between the air mode and the CO2 mode leads a small before stall, and become smaller as Mach number increases. It is
difference in lift coefficients at high angles of attack. In particular, the considered that the difference in high angles of attack where the
transition point exists forward of the maximum thickness point above mixing of the separated shear layer is enhanced by a strong three-
α ! 8 deg at M ! 0.20, and then the mixing of the separated shear dimensionality of the flow is due to the difference in momentum
layer is enhanced in the large area on the wing. As a result, the exchange by Reynolds stress between the air mode and the CO2
difference in lift is observed at relatively low angles of attack (around mode. However, these differences in lift coefficients are small.
α ! 8 deg). Meanwhile, the effect of compressibility delays the Therefore, it can be concluded that specific heat ratio has negligible
transition at higher Mach number. In addition to stabilization of the effect on aerodynamic performance even when the Reynolds and
separated shear layer, an influence region of momentum exchange by Mach numbers change.
the flow mixing also decreases on the wing and consequently leads Further discussion associated with the flowfield change between
lower effects of specific heat ratio on aerodynamic performance at the air and CO2 modes requires detailed pressure distribution
higher Mach number. measurement. However, to date, the PSP technique has not yet been
established in the CO2 mode. In addition, unsteady flow phenomena
VII. Conclusions such as vortex shedding at high angles of attack probably occur on the
model surface. To clarify these flow phenomena, the PSP measure-
Aerodynamic characteristics and pressure distributions of a flat ment techniques for the CO2 case and unsteady measurements under
plate with 5% thickness and a NACA 0012-34 airfoil were the low-pressure condition are under development.
investigated in the Mars Wind Tunnel to evaluate the similarity rule
under low-Reynolds-number and high-subsonic-flow conditions.
For the flat plate, the Reynolds number affects the lift-curve slope
and drag characteristics, whereas the Mach number and specific Acknowledgments
heat ratio have almost no impact on aerodynamic performance. A We would like to express our sincere thanks to S. Obayashi and T.
leading-edge separation bubble with a distinguishing low-pressure Ogawa of Tohoku University for their continuing support to our
region was captured by PSP measurement. In particular, laminar study. We would also like to thank M. Okamoto of Kanazawa
reattachment without a laminar-to-turbulent transition was observed Institute of Technology for valuable discussion on low-Reynolds-
only at α ! 0 deg for Re ! 4.9 × 103, indicating that a critical number airfoils. We thank also T. Ono, K. Nose, and S. Ida, who
region where the reattachment state changes exists between Re ! graduated from Tohoku University, for this research. This research
4.9 × 103 and Re ! 1.1 × 104 . Compressibility has an effect of was supported by a Grant-in-Aid for Scientific Research
stabilizing the separated shear layer and delaying transition to (KAKENHI: 2056081) and the Japan Aerospace Exploration
turbulence and reattachment. However, this effect of the Agency.
1654 ANYOJI ET AL.

References Paints to Low Pressure Range,” Journal of Thermophysics and Heat


Transfer, Vol. 19, No. 1, 2005, pp. 9–16.
[1] Mueller, T. J., Kellogg, J. C., Ifju, P. G., and Shkarayev, S. V., doi:10.2514/1.5047
Introduction to the Design of Fixed-Wing Micro Air Vehicles: Including [20] Mori, H., Niimi, T., Hirako, M., and Uenishi, H., “Pressure Sensitive
Three Case Studies, AIAA, Reston, VA, 2007, pp. 1–107. Paint Suitable to High Knudsen Number Regime,” Measurement
[2] Jacob, J. D., and Smith, S. W., “Design of HALE Aircraft Using Science and Technology, Vol. 17, No. 6, 2006, pp. 1242–1246.
Inflatable Wings,” 46th AIAA Aerospace Sciences Meeting and Exhibit, doi:10.1088/0957-0233/17/6/S02
AIAA Paper 2008-0167, Jan. 2008.
[21] Ono, T., “Development of Pressure-Sensitive Paint Technique for
[3] Platzer, M. F., Jones, K. D., Young, J., and Lai, J. C. S., “Flapping Wing Surface Pressure Measurement in a Mars Wind tunnel,” Proceedings of
Aerodynamics: Progress and Challenges,” AIAA Journal, Vol. 46, No. 9, the 14th International Symposium on Flow Visualization, Korean Soc.
2008, pp. 2136–2149. of Visualization, No. 107, ISFV12-4D-5, 2010.
doi:10.2514/1.29263 [22] Rae, W. H., Jr., and Pope, A., Low-Speed Wind Tunnel Testing, 2nd ed.,
[4] McMasters, J. H., and Henderson, M. L., “Low-Speed Single-Element Wiley, New York, 1984, p. 176.
Airfoil Synthesis,” Technical Soaring, Vol. 6, No. 2, 1980, pp. 1–21.
[23] Mueller, T. J., and Burns, T. F., “Experimental Studies of the Eppler 61
[5] Laitone, E. V., “Aerodynamic Lift at Reynolds Numbers Below
Airfoil at Low Reynolds Numbers,” 20th Aerospace Sciences Meeting,
7 × 104,” AIAA Journal, Vol. 34, No. 9, 1996, pp. 1941–1942. AIAA Paper 1982-0598, Jan. 1982.
doi:10.2514/3.13329 [24] Gerakopulos, R., Boutilier, M. S. H., and Yarusevych, S., “Aerodynamic
[6] Schmitz, F. W., “Aerodyamics of the Model Airplane Part 1,” Redstone Characterization of a NACA 0018 Airfoil at Low Reynolds Numbers,”
Scientific Information Center Rept. 721, Alabama, 1967. 40th Fluid Dynamics Conference and Exhibit, AIAA Paper 2010-4629,
[7] Schmitz, F. W., “The Aerodyamics of Small Reynolds Number,” NASA June–July 2010.
TM-51, 1980.
[25] Okamoto, M., and Azuma, A., “Aerodynamic Characteristics at Low
[8] Ohtake, T., Nakae, Y., and Motohashi, T., “Nonlinearity of the
Downloaded by KYUSHU UNIVERSITY OOA on July 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/1.J053468

Reynolds Numbers for Wings of Various Planforms,” AIAA Journal,


Aerodynamic Characteristics of NACA0012 Aerofoil at Low Reynolds Vol. 49, No. 6, 2011, pp. 1135–1150.
Numbers,” Journal of the Japan Society for Aeronautical and Space doi:10.2514/1.J050071
Sciences, Vol. 55, No. 644, 2007, pp. 439–445. [26] Garmann, D. J., and Visbal, M. R., “Dynamics of Revolving Wings for
doi:10.2322/jjsass.55.439 Various Aspect Ratios,” Journal of Fluid Mechanics, Vol. 748,
[9] Guynn, M. D., Croom, M. A., Smith, S. C., Parks, R. W., and Gelhausen,
June 2014, pp. 932–956.
P. A., “Evolution of a Mars Airplane Concept for the ARES Mars Scout
doi:10.1017/jfm.2014.212
Mission,” 2nd AIAA “Unmanned Unlimited” Conference and [27] Sasaki, K., and Kiya, M., “Three-Dimensional Vortex Structure in a
Workshop & Exhibit, AIAA Paper 2003-6578, Sept. 2003. Leading-Edge Separation Bubble at Moderate Reynolds Number,”
[10] Braun, R. D., and Spencer, D. A., “Design of the ARES Mars Airplane Journal of Fluids Engineering, Vol. 113, No. 9, 1991, pp. 405–410.
and Mission Architecture,” Journal of Spacecraft and Rockets, Vol. 43, doi:10.1115/1.2909510
No. 5, 2006, pp. 1026–1034.
[28] Tani, I., Iuchi, M., and Komoda, H., “Experimental Investigation of
doi:10.2514/1.17956
Flow Separation Associated with a Step or a Groove,” Aeronautical
[11] Tanaka, Y., Okabe, Y., Suzuki, H., Nakamura, K., Kubo, D., Tokuhiro, Research Inst. Rept. 364, Univ. of Tokyo, Tokyo, 1961.
M., and Rinoie, K., “Conceptual Design of Mars Airplane for [29] Alam, M. M., Zhou, Y., Yang, H. X., Guo, H., and Mi, J., “The Ultra-
Geographical Exploration,” Journal of the Japan Society for Low Reynolds Number Airfoil Wake,” Experiments in Fluids, Vol. 48,
Aeronautical and Space Sciences, Vol. 54, No. 624, 2006, pp. 24–26. No. 1, 2010, pp. 81–103.
[12] Oyama, A., and Fuji, K., “A Study on Airfoil Design for Future Mars doi:10.1007/s00348-009-0713-7
Airplane,” 44th AIAA Aerospace Sciences Meeting and Exhibit, AIAA
[30] Zhou, Y., Alam, M. M., Yang, H. X., Guo, H., and Wood, D. H., “Fluid
Paper 2006-1484, Jan. 2006.
Forces on a Very Low Reynolds Number Airfoil and Their Prediction,”
[13] Drela, M., “Transonic Low-Reynolds Number Airfoils,” Journal of International Journal of Heat and Fluid Flow, Vol. 32, No. 1, 2011,
Aircraft, Vol. 29, No. 6, 1992, pp. 1106–1113. pp. 329–339.
doi:10.2514/3.46292 doi:10.1016/j.ijheatfluidflow.2010.07.008
[14] Anyoji, M., Nose, K., Ida, S., Numata, D., Nagai, H., and Asai, K., [31] Azuma, A., Okamoto, M., and Yasuda, K., “Aerodynamic Character-
“Development of a Low-Density Wind Tunnel for Simulating Martian
istics of Wing at Low Reynolds Number,” Fixed and Flapping Wing
Atmospheric Flight,” Transactions of the Japan Society for Aerospace
Aerodynamics for Micro Air Vehicle Applications, edited by Mueller, T.
for Aerospace and Space Sciences, Vol. 9, 2011, pp. 21–27. J., Vol. 195, Progress in Astronautics and Aeronautics, AIAA, Reston,
[15] Anyoji, M., Ida, S., Nose, K., Numata, D., Nagai, H., and Asai, K., VA, 2001, pp. 341–398.
“Characteristics of the Mars Wind Tunnel at Tohoku University in CO2 [32] Suwa, T., Nose, K., Numata, D., Nagai, H., and Asai, K.,
Operation Mode,” 48th AIAA Aerospace Sciences Meeting, AIAA “Compressibility Effects on Airfoil Aerodynamics at Low Reynolds
Paper 2010-1490, Jan. 2010. Number,” 30th AIAA Applied Aerodynamics Conference, AIAA Paper
[16] Anyoji, M., Nose, K., Ida, S., Numata, D., Nagai, H., and Asai, K., “Low
2012-3029, June 2012.
Reynolds Number Airfoil Testing in a Mars Wind Tunnel,” 40th Fluid
[33] Anyoji, M., Nonomura, T., Aono, H., Oyama, A., Fujii, K., Nagai, H.,
Dynamics Conference and Exhibit, AIAA Paper 2010-4627, June– and Asai, K., “Computational and Experimental Analysis of a High-
July 2010. Performance Airfoil Under Low-Reynolds-Number Flow Condition,”
[17] Anyoji, M., Nose, K., Ida, S., Numata, D., Nagai, H., and Asai, K., Journal of Aircraft, 2014 (accepted for publication).
“Aerodynamic Measurements in the Mars Wind Tunnel at Tohoku doi:10.2514/1.C032553
University,” 49th AIAA Aerospace Sciences Meeting, AIAA Paper
2011-0852, Jan. 2011. F. Coton
[18] Barlow, J. B., Rae, W. H., Jr., and Pope, A., “Low-Speed Wind Tunnel Associate Editor
Testing,” 3rd ed., Wiley, New York, 1999, p. 375.
[19] Niimi, T., Niimi, T., Yoshida, M., Kondo, M., Oshima, Y., Mori, H.,
Egami, , Asai, K., and Nishide, H., “Application of Pressure-Sensitive

View publication stats

You might also like