You are on page 1of 16

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

net/publication/306033282

Lift and Drag Measurements of a Gull-Wing Configuration Aircraft

Conference Paper · January 2015


DOI: 10.2514/6.2015-0027

CITATIONS READS
5 2,559

2 authors:

Tyler Davis Geoffrey Spedding


Calspan Corporation University of Southern California
4 PUBLICATIONS   12 CITATIONS    150 PUBLICATIONS   4,907 CITATIONS   

SEE PROFILE SEE PROFILE

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

Bio(fluid)mechanics of bat flight View project

Exulans Project View project

All content following this page was uploaded by Geoffrey Spedding on 14 August 2017.

The user has requested enhancement of the downloaded file.


Lift and Drag Measurements of a Gull-Wing Configuration
Aircraft

T.W. Davis1 and G.R. Spedding2


University of Southern California, Los Angeles, CA, 90089

The tube-and-wing aircraft configuration has remained unchanged since the start of
commercial aviation. The state of minimum induced drag is comprised of an elliptically
loaded wing, which has a constant downwash profile. Interrupting this wing with a central,
payload-carrying body costs downwash, but is essential in fulfilling practical flight
objectives. Initial drag and lift estimates, streamwise velocity profiles, as well as sectional
and quasi-3D parasite drag estimates have been produced using force balance and Particle-
Imaging-Velocimetry (PIV) measurement techniques on a novel aircraft configuration,
called the Gull-Wing Configuration. The results here indicate it is possible to restore an
estimated 59% of the lost downwash while incurring no measurable increase in drag because
of induced drag reductions, despite increases in parasite drag.

Nomenclature
AR = wing aspect ratio
b = wingspan [cm]
c = chord length [cm]
d = fuselage diameter [cm]
d’ = distance from observation plane centerline to GW model centerline [cm]
D = drag force [N]
D’ = drag force per unit span [N/m]
Di = induced drag [N]
e = Oswald efficiency factor
k = induced drag factor
l = body length [cm]
lo = distance from wing chord-line to model body centerline [cm]
L = lift force [N]
L/D = lift to drag ratio
q = dynamic pressure [Pa]
Q = root mean square (RMS) of velocity fluctuations [m/s]
Rec = chord-based Reynolds number
S = wing planform area [ m2 ]
U = nominal mean freestream streamwise velocity [m/s]
Uref = empirical mean streamwise velocity [m/s]
(u,v,w) = streamwise, spanwise, and cross-stream velocity components [m/s]
wi = induced velocity, or downwash, due to wing lift [m/s]
(x,y,z) = streamwise, spanwise, and cross-stream coordinates [m]
α = wing set angle [˚]
 = angle between model aft-body and model fore-body centerline [˚]
t = time between laser pulses [μs]

1
Graduate Research Assistant, Department of Aerospace and Mechanical Engineering, tylerdav@usc.edu, Student
Member AIAA.
2
Professor, Department of Aerospace and Mechanical Engineering, geoff@usc.edu, Member AIAA.
1
American Institute of Aeronautics and Astronautics
ε = fractional increase in induced drag for non-elliptically loaded wing
ν = kinematic viscosity [m2/s]
 = density of air [kg/m3]
θ = angle between freestream flow and model centerline [˚]
Θ = momentum thickness [cm]
Cd = sectional (2D) drag coefficient
CD = total 3D drag coefficient
Cd,p = 2D parasite drag coefficient
CD,i = 3D induced drag coefficient
CD,p = quasi-3D parasite drag coefficient
Cl = sectional or 2D lift coefficient
CL = 3D lift coefficient

I. Introduction

T hough the optimum circulation distribution on a planar wing has an elliptic spanwise variation, dropping
gradually from its maximum at the centerplane (y = 0), the Current Dominant Configuration (CDC), which is
almost universally adopted in passenger and cargo transports, does not have such a circulation distribution [1,2].
The wing in the CDC is interrupted at some point by a central fuselage, and though there are a number of careful
designs merging the wing into the body to avoid discontinuous jumps, the basic design of the CDC is unaltered. The
idea that there could be aerodynamically more efficient designs has led to repeated configuration studies and much
work on certain aspects, such as blended wing bodies (BWB) [3].
Based on a reconsideration of some of the guiding principles and possibilities should passive pitch stability be
achieved through wing shape change alone, it has been proposed [4] that a new aircraft configuration with a shorter,
wider body and a control surface attached to a deflected aft-body can change the circulation about an entire wing-
body assembly so as to restore the uniform downwash profile (and, by implication, the elliptical loading
distribution). Because the aft control surface changes the location of the rear stagnation point, it is called a Kutta-
Edge (KE). Through initial Particle Imaging Velocimetry (PIV) measurements, Ref. [4] showed that this novel
aircraft configuration can be more aerodynamically efficient than the CDC, while still maintaining a central and
dedicated payload-carrying body. Perhaps not coincidentally, this independent design resembles that of a sea-bird,
so it is termed the Gull-Wing (GW) configuration.
Though these flow-field measurements were encouraging, direct and confirmatory force balance measurements
were not possible due to the weight of the model. The primary focus of this paper is on making direct drag and lift
measurements of a lightweight GW configuration aircraft as functions of only two geometric parameters.
Instantaneous and time-averaged velocity fields are also measured using PIV, where the momentum thickness in the
wake of the GW model is calculated, leading to sectional and quasi-3D parasite drag estimates. Force balance and
PIV measurement techniques are combined to quantify the aerodynamic performance benefits, and determine likely
sources of drag and lift variation by breaking drag into induced and parasite components.

II. Materials and Methods

A. Dryden Wind Tunnel


Experiments were performed in the Dryden Wind Tunnel at USC. It is a closed-loop, low turbulence (Q/U < 0.03
%) wind tunnel with an octagonal cross-section, with 1.36 m between opposing faces. The mean flow speed for
these tests was set at a constant U = 21 m/s. The resulting chord-based Reynolds number Rec = Uc/ν = 1.04  105 .
Total drag and lift were estimated directly through force balance measurements and the parasite drag was estimated
indirectly through PIV measurements. Figure 1 shows the experimental set-up.

B. Gull-Wing Model
The GW model is made from two sections, the fore- and aft-bodies, which can be arranged with only two
adjustable parameters, the aft-body deflection angle, δ, and the angle between the freestream flow and the GW body
centerline, θ. The aft-body is deflected at five different angles, δ = 0⁰, 2⁰, 4⁰, 6⁰, and 8⁰, using five acrylic wedges.
Molding clay is applied around the circumference of the wedge to smooth the transition from the fore-body to the
aft-body.
2
American Institute of Aeronautics and Astronautics
KE and NKE denote arrangements with and without the Kutta-Edge tail plate, respectively. The difference can
be seen in Fig. 2(b) and (c). The GW model is 28.55 cm in length in the NKE arrangement with a maximum body
diameter d = 6 cm. The body fineness ratio l/d = 4.75. The GW model uses a NACA 0012 airfoil to make up its
rectangular, untwisted wing. The wingspan b = 50.0 cm, and the chord-length c = 7.5 cm. The wing is fixed at a set
angle of attack α = 5.0⁰ ± 0.5⁰, and the chord-line of the wing is displaced from the centerline of the body by lo =
1.87 cm.
A solid steel rod, 0.7 cm in diameter, traverses 80% of the wingspan and is fixed inside the wing at the ¼-chord.
This rod protrudes out of one wing-tip and is inserted into a sting (Fig. 1). The model is mounted at the wing-tip,
instead of the belly of the GW model to minimize disturbances to the wing-fuselage-junction. This shrouded sting
projects through the wind tunnel floor onto a force balance.
The ideas behind the GW configuration are essentially inviscid, or Re-independent, and so wind tunnel tests at
lower Re must be carefully justified and controlled. Boundary-layer trips are placed on both body and wings to
avoid Re-dependent variations due to flow separation and possible reattachment. In particular, junction flows may
be adversely affected by growing and/or separating boundary layers at Re c = 105. If the benefits of the GW
configuration can be measured at Rec = 105, there is little reason to believe the same behavior would not be present
at larger Rec.

Camera and Optics

GW Prototype
x z
y Laser Sheet Envelope

Sting
Force Balance
Rotary Table
Figure 1. Dryden Wind Tunnel test section showing the GW model,
force balance, laser sheet, and camera.

The dynamic pressure of the flow is

1 (1)
q U 2 .
2

Depending on the precise air temperature, the product qS is approximately 10 N, where S is the wing planform area,
S = bc. The drag and lift coefficients for a finite wing are

D (2)
CD 
qS

L (3)
CL  .
qS

3
American Institute of Aeronautics and Astronautics
From Ref. [5], at α = 5˚ the NACA 0012 has a Cl = 0.55 and a Cd = 0.0075 at Rec = 3 106 . The smallest expected
non-zero force will be D = qSCd = 75 mN, when L = 5.5 N.

C. Direct Drag and Lift Measurements


Time-averaged drag and lift measurements were made with a custom force balance. The force balance has a
cruciform shape, with four strain gauges per arm, arranged as four full Wheatstone bridges for high sensitivity (ΔD,
ΔL < 10 mN) [6]. Calibrations were conducted over 78-784 mN in drag, and 0-3492 mN in lift, to generate a 3  4
calibration matrix that includes leading order cross-terms of L(D) and D(L). The deflection angles tested were  =
0˚, 2 ˚, 4 ˚, 6˚, and 8˚ in both the KE and NKE arrangements for θ = -5˚ to θ = 1˚.

α
δ
θ
U
(a) FB AW AB KE

KE NKE

(b) (c)
Figure 2. (a) side view of the GW model, showing θ and δ. FB = fore-body, AW = acrylic
wedge, and AB = aft-body. (b) GW model in the KE arrangement (c) GW model in the
NKE arrangement.

Using the direct drag and lift measurements, CD and CL can be estimated. The induced cross-stream velocity, or
downwash, wi(z), can be estimated from

C L AR (4)
wi  .
U

The total drag coefficient of the model is

C D  C D,i  C D,p , (5)

where CD,i is the induced drag coefficient, commonly written

4
American Institute of Aeronautics and Astronautics
CL 2 (6)
C D,i  .
eAR

CL is the total lift coefficient and e is the Oswald efficiency factor, defined as

1 (7)
e .
1    kAR

ε is the fractional increase in induced drag beyond the optimum, elliptically loaded wing and k is found empirically
from the slope of CD vs. CL2 [7-9]. AR is the aspect ratio defined as AR = b2/S. AR = 6.67 for the wing used in the
GW model. CD,p is the parasite drag coefficient, which includes all forms of drag independent of lift generation [8].
Force balance measurements yield an estimate of CD and CL directly.
The parasite drag coefficient may also be measured indirectly through the wake momentum thickness. However,
on an aircraft model with a number of separate lift and drag generating surfaces, a net lift of zero does not guarantee
that the parts themselves generate zero lift, and so does not guarantee that their induced drag will be zero either,
except in the curve fitting sense of simple equations such as Eq. (5). Thus, a direct CD,p comparison between force
balance and PIV measurements cannot be made.

D. Indirect Drag Measurements: Estimate of CD,p


1. Theoretical Background
The parasite drag coefficient of the GW model can be estimated from the model’s wake defect (Fig. 3). Using

b
a
z
x
const

c
d
Figure 3. Comparison of two different flow regions used for deriving the
momentum thickness and estimating Cd,p.

the integral form of the Navier-Stokes equations and Conservation of Mass through the streamtube (a-b-c-d) in Fig.
3, one may write

d u u 
2 (8)
D'  U 2
b
 2  2 
 U  U 
dz ,


for a two-dimensional, non-turbulent, steady flow with no residual pressure gradients. Equation (8) can be expressed
in terms of the momentum thickness,
d u u  
(9)
2


Θ   2   2  dz ,
b U
  U  

5
American Institute of Aeronautics and Astronautics
as

D'  U 2 . (10)

From Eq. (2), the sectional drag coefficient is

D' (11)
Cd  ,
qS

and substituting Eqs. (9), (1), and (10) into Eq. (11), with S = bc yields

2 (12)
C d,p  .
c

u2(z) is measured in the (x,z) plane, at multiple spanwise (y) locations. Thus, a 3D wake defect volume was in
effect acquired to estimate a quasi-3D parasite drag coefficient, CD,p. One method of estimating CD,p from Cd,p is
adding up Cd,p found at each spanwise location, multiplying by the increment used for changing the spanwise
location, and then dividing by the total span over which the measurements were made. From the GW model
symmetry, the quasi-3D parasite drag coefficient for the entire GW model is

 b
2  (13)
2 
C D,p
b 
 C d,p, y 0  2  ( y )dy .
 1 

2. Equipment
The (u,w) velocity components in 2D planes (x,z) were measured as shown in Fig. 4. The PIV system used a
Nd:YAG laser, and a LaVision Imager Pro X 2M camera (1600  1200  14 bit pix) with a 70:210 mm focal length
Nikon lens. The laser-sheet generating optics and the camera reside on independent 3-axis traverses, controlled
using one stepper motor per axis. To build a wake defect volume, the camera and optics traversed the spanwise (y)
direction in unison.

3. Processing and estimating CD,p


36 spanwise locations, shown in Fig. 4(b), were sampled from y/b = -0.1 to y/b = 0.6, in Δy/b = 0.02 increments,
where model lift was estimated to be approximately zero in initial tests. Only the y/b = 0 to y/b = 0.5 spanwise
locations were used in estimating CD,p. The time between laser pulses was δt = 80 μs. Similar to the force balance
measurements, all five  angles were tested ( = 0˚, 2 ˚, 4 ˚, 6˚, and 8˚) in the KE and NKE model arrangements, but
at θ = -4 ˚ and θ = -3˚ for  = 2˚, 4˚, 6˚, and 8˚, and at θ = -3 ˚ for  = 0˚. CD corresponding to the zero-CL value was
found from the force balance results, and then used to find the corresponding θ in the CD(θ) plot. This was done for
each , yielding the θ values used for the PIV measurements. For all aft-body deflections except δ = 0˚, θ was
between θ = -4˚ and θ = -3˚. Thus, both θ were tested during the PIV measurements. The corresponding θ for the δ =
0˚ deflection was θ = -3˚. θ corresponding to the zero-lift drag was used because if CD,i is not zero here, it is likely
that it will be a non-zero minimum.
200 image pairs were acquired at each spanwise location at 9.6 Hz. An in-house Correlation Imaging
Velocimetry (CIV) program was used to generate and post-process the velocity fields that are then reinterpolated
and smoothed using a 2D cubic spline interpolator, explained at length in Ref. [11]. An average over x/c = 0.073 to
x/c = 0.133 was taken to yield a robust estimate of the raw wake velocity profile, without incurring further
momentum loss. (See Results, Section B.1. for further details.)
Wind tunnel flows over a finite model do not produce ideal velocity distributions such as shown in Fig. 3. The
velocity defect is super-imposed on some non-uniform background. A fifth order polynomial was fit to the baseline
of this time-averaged, raw streamwise wake velocity profile, u(z), to define the mean baseline velocity profile, UB(z).
The wake behind the GW model is superimposed on top of UB(z). The baseline corrected velocity is u*,

6
American Institute of Aeronautics and Astronautics
u*  U B ( z)  u( z) , (14)

and is substituted for u2 in Eq. (9). u*(z) are normalized by Uref, an empirical mean freestream velocity in the x
direction, estimated from the upper left-hand corner of the time-averaged velocity field in a sub-window far from the
model.

ALL DIMENSIONS ARE IN cm


z
¼-Chord Midline x
Body Midline d' = 0.75
Observation Volume Observation
(10.6 x 8.0 x 25.0) Plane Midline
(a)
y
y/b = 0.6
Wing Tip
(y/b = 0.5)
c = 7.5
Δ y/b = 0.02

DPIV
sampling
locations

Body Midline
x
(y/b = 0)
y/b = -0.1
(b)
Figure 4. (a) observation volume and (b) spanwise locations for the PIV measurements.

III. Results

A. Force Balance Drag and Lift Measurements


CD(δ) is shown in Fig. 5 for θ = 0˚ and θ = -2˚. θ = 0˚ is similar to a cruise condition, and neither the KE nor aft-
body deflection appear to aid in drag reduction. The increase in drag coefficient from ˚ to ˚ is 9% for the
KE arrangement, but only 5% for the NKE arrangement.
The data for θ = -2˚ show a decrease in drag coefficient for ˚ or ˚, compared with the base ˚.
Moreover, CD is the same for ˚ as ˚. These points are not seen in NKE measurements. The KE is therefore
responsible for the decrease in CD for ˚ or ˚ and for the comparatively low CD at˚.
CL() is shown in Fig. 6. There is a 23% increase in CL when the KE deflection angle increases from ˚ to ˚,
and only a 5% increase in CL for the same variation in NKE arrangement. Figure 6 shows that aft-body deflection
alone can improve CL, but it does so much less than with the KE.

7
American Institute of Aeronautics and Astronautics
0.036

0.035

0.034

0.033

CD
0.032

0.031

0.03
KE
NKE
0.029
0 2 4 6 8

δ [⁰]
(a)

0.032

0.0315

0.031

0.0305

0.03
CD 0.0295

0.029

0.0285

0.028 KE
NKE
0.0275
0 2 4 6 8

δ [⁰]
(b)

Figure 5. Total CD(δ) at θ = 0˚ in plot (a) and θ = -2˚ in plot (b). The error bars come from maximum
differences in any series of experiments at the same conditions.

8
American Institute of Aeronautics and Astronautics
0.37

0.35

0.33

0.31
CL
0.29

0.27
KE
NKE
0.25
0 2 4 6 8

δ [⁰]
Figure 6. CL( at θ = 0⁰.

Equation (4) relates wi and CL and one may compare CL with values expected (and measured in these
experiments) for a finite wing alone at the same Rec. CL in Fig. 6 is lower than for a clean wing because a large
central fuselage disrupts the circulation distribution, and the changes in CL may also be expressed as changes in wi,
perhaps a more direct way of envisioning the role of the KE.
The lost downwash fraction (compared with a clean wing) for ˚ (wi = -0.27 m/s) is 34%, and the lost
downwash for ˚ (wi = -0.35 m/s) is 14%, in the KE arrangement. So, 59%, or (34-14)/34, of the lost downwash
is restored when ˚, where total drag increases by 1.6%. Without the KE, only 27% of the lost downwash can be
restored at best (˚), where the drag increase is 6%. Figure 7 shows these changes, where Δwi is the percentage
of downwash lost as compared with the NACA 0012 wing only case, and ΔCD is the increase in drag above the
˚, NKE case. Figure 7 shows how the deflected KE can restore the central downwash, while incurring little or
negative increase in drag.
L/D(is shown in Fig. 8. Using only the average values, the optimum aft-body deflection angle for KE is
˚, and the optimum for NKE has a broader plateau between ˚ and ˚. There is a significant difference
between the base (˚) deflection and the estimated optimal values. The increases in L/D at the optimum  for the
KE arrangement and the NKE arrangement are 18% and 7%, respectively. The shape of the L/D curves is strongly
influenced by the CL curves because the CL values are an order of magnitude larger than CD. Figure 8 also shows
that it is the combination of the aft-body deflection and KE that provide aerodynamic performance benefits. Notice
the higher average L/D for ˚ and ˚ in the NKE arrangement as compared with the KE arrangement,
presumably because the extra friction drag incurred by the tail area first needs to be paid.

B. PIV Parasite Drag Estimates


1. Streamwise sampling location
The formulation of Eq. (8) presumes that pressure gradients are negligible, a condition typically found after x/c =
2-3 [10]. Figure 9 shows the calculated Θ/c(x/c) for NKE. Θ/c continually decreases with increasing x/c. The flow
over the wing-body junction is turbulent and losses occur before the measurement plane. In the absence of a better
alternative, Θ/c is as far upstream as possible, but is likely to be an underestimate of the true value.

9
American Institute of Aeronautics and Astronautics
40 KE NKE

35

30

25

Δwi 20
[%]
15
δ = 0⁰
10 δ = 2⁰
δ = 4⁰
5 δ = 6⁰
δ = 8⁰
0
-7 -5 -3 -1 1 3 5 7
ΔCD [%]
Figure 7. Δwi(ΔCD) for all tested δ, in the KE and NKE arrangements at θ = 0˚.

10.05

9.75

9.45

L/D 9.15

8.85

8.55 KE
NKE
8.25
0 2 4 6 8
δ [˚]
Figure 8. L/D(at θ = 0˚.
10
American Institute of Aeronautics and Astronautics
0.1
0.09
0.08
0.07
0.06
Θ/c 0.05
0.04
0.03
0.02
0.01
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4
x/c
Figure 9. Θ/c(x/c) for 2˚, NKE at y/b = 0 and θ = -4˚.

0.4 0.4

0.2 0.2

0 0

-0.2 -0.2
z/c
z/c

-0.4 -0.4

-0.6 -0.6

-0.8 -0.8

-1 -1
0.7 0.8 0.9 1 1.1 -0.1 0 0.1 0.2 0.3 0.4
u u$
Uref Uref
(a) (b)
Figure 10. (a) u(z) is the solid line with square symbols, and UB(z) is the dashed line. (b) u*(z) is the solid
line with square symbols. These data are from y/b = 0, x/c = 0.088, and δ = 6˚, KE. y/b = 0 and θ = -4˚.

2. Processing streamwise velocity profiles


The raw streamwise wake velocity profile u(z) is plotted in Fig. 10(a), where the polynomial fit mean outer-flow
UB(z) is shown with the dashed line. Fig. 10(b) shows the corrected u*(z).
The double-lobe form of u*(z) is an early wake remnant of the upper and lower boundary layers shed into the
wake. Figure 11(a) shows a small δ = 2˚ aft-body deflection for both KE and NKE. Figure 11(b) shows the higher δ
= 6˚ aft-body deflection for both KE and NKE. The double-lobe structure is more pronounced with KE as it
separates the upper and lower free shear layers originating from separation before the end of the body. We may note
also how the mean defect profiles are deflected lower in z at higher , consistent with the inferred and measured
higher lift due to the KE.
Figure 12 shows u*(z) for δ = 0˚, 2˚, and 6˚, all with the KE. From Fig. 5(b), it is expected that either the parasite
drag, the induced drag, or both, will decrease when δ increases from 0˚ to 2˚. Quick inspection of the profiles at
different  indicate that CD,p does not decrease, but increases. Already we see an implication that KE decreases the
induced drag component.

3. Estimating Cd,p and CD,p


Figure 13 shows the spanwise variation of Cd,p for δ = 0˚, 2˚, and 6˚ aft-body deflections. The Cd,p values are
comparable to reference values for the NACA 0009 by y/b = 0.1 [12]. Since CD,p in Eq. (13) is dominated by the

11
American Institute of Aeronautics and Astronautics
momentum defect for -0.02 ≤ y/b ≤ 0.02, or the middle 4% of the model, decreasing the step size Δy would help
greatly in decreasing ΔCD,p.
The quasi-3D parasite drag coefficient CD,p(δ) is shown in Fig. 14. On average, ΔCD,p are 10% of the average
CD,p values for all δ. In Fig. 14, the KE increases CD,p for all non-zero δ. This is expected as the KE may be thought
of as a control surface whose primary purpose is to decrease induced drag. The force balance data in Fig. 14, shown
with the square symbols, show the total drag coefficient, CD. The difference between FB and PIV is equal to CD,i.

0.4

0.2

-0.2
z/c

-0.4

-0.6

-0.8

-1
-0.1 0 0.1 0.2 0.3 0.4
$
u
Uref
(a)

0.4

0.2

-0.2
z/c

-0.4

-0.6

-0.8

-1
-0.1 0 0.1 0.2 0.3 0.4
$
u
Uref
(b)

Figure 11. Time-averaged, mean outer-flow subtracted velocity profiles for δ = 2˚ (a), and δ = 6˚ (b). The
solid line represents the KE arrangement, the dotted line represents the NKE arrangement, circle symbols
are for δ = 2˚, and triangle symbols are for δ = 6˚. y/b = 0, x/c = 0.088, and θ = -3˚.

12
American Institute of Aeronautics and Astronautics
0.4

0.2

-0.2
z/c

-0.4

-0.6

-0.8

-1
-0.1 0 0.1 0.2 0.3 0.4
$
u
Uref
Figure 12. Time-averaged, baseline-corrected streamwise velocity profiles for δ = 0˚ (solid black line,
square symbols), δ = 2˚ (dashed green line, circle symbols), and δ = 6˚ (dotted blue line, triangle
symbols), KE. y/b = 0, x/c = 0.088, and θ = -3˚.
0.09
δ=0˚, KE
0.08
δ=2˚, KE
0.07 δ=6˚, KE
NACA0009
0.06

0.05
Cd,p 0.04

0.03

0.02

0.01

0
-0.1 0 0.1 0.2 0.3 0.4 0.5 0.6
y/b
Figure 13. Cd,p (y/b). Cd,p = 0.0106 for a NACA0009 at Rec =105 is shown with the dotted line. (θ = -3˚).
Symbol sizes scale with uncertainty.

13
American Institute of Aeronautics and Astronautics
FB, KE
0.034 FB, NKE
PIV, KE
CD PIV, NKE
0.031

CD,p0.029
0.026

0.024

0.021
0 2 4 6 8
δ [˚]
Figure 14. CD,p(δ) from PIV measurements, and CD(δ) from force balance (FB) measurements. Both the KE
and NKE arrangements are shown.

IV. Summary and Conclusions


The Gull-Wing configuration has been proposed as a means of reducing drag behind a wing-body combination
where neither wing nor body designs are compromised by being forced to merge or blend into one another. PIV-
based investigations [4] have shown that the spanwise distribution of induced downwash was strongly affected by
the KE, and that the magnitude of the deflection angle, , could fill in the missing downwash in the central fuselage
wake in the predicted fashion. The KE could be used either to reduce Di or to increase lift (regardless of cost).
Here, combined force balance and PIV measurements show that the deflected KE increases the parasite (zero-
lift) drag, but that the total drag increases by less, because Di has decreased. Therefore the overall flight performance
as measured by the usual L/D figure of merit, can increase by 18%. In times of slender profit margins and high
relative fuel costs, such performance gains are not negligible. They are also accompanied by a different and
uncompromised fuselage design which can be further tailored for optimal packing efficiency and/or reduced drag.
The results described here come from tests at low Rec and despite the presence of trip-strips, one may still
wonder how fuselage boundary layers can most efficiently interact with the tail geometry. This is a topic for future
study.

Acknowledgments
This work has been supported through internal grants at USC. T.W. Davis dedicates this paper to B.K. Dovey.

References
1Munk, M. M., “The Minimum Induced Drag of Aerofoils,” NACA TR-121, 1923.
2Jones, R.T., Wing Theory, Princeton Univ. Press, Princeton, NJ, 1980, Chap. 7, pp. 105-128.
3Liebeck, R.H., “Design of the BWB Subsonic Transport”, AIAA-2002-0002, 40th AIAA Aerospace Sciences Meeting and

Exhibit, January 14-17, 2002, Reno, NV.


4Huyssen, R.J., Spedding, G.R., Mathews, E.H., and Liebenberg, L., “Wing-body circulation control by means of a fuselage

trailing edge,” Journal of Aircraft, Vol. 49, No. 5, 2012, pp. 1279-1289.
5Abbott, I.H., and Von Doenhoff, A.E., Theory of Wing Sections, Dover ed., McGraw-Hill Book Company, Inc., New York,

1959, App. IV, pp.460-461.

14
American Institute of Aeronautics and Astronautics
6Zabat, M., Farascaroli, S., Browand, F., Nestlerode, M., and Baez, J. “Drag Measurements on a Platoon of Vehicles.” California
Partners for Advanced Transit and Highways(PATH) Program, Institute of Transportation Studies, University of California
Berkeley, 1994, p. 18-31.
7Yechout, T.R., Introduction to Aircraft and Flight Mechanics, AIAA Education Series, Reston, 2003, Chap. 1, pp. 50-51.
8McCormick, B.W., Aerodynamics, Aeronautics, and Flight Mechanics, 2nd Ed., Wiley-India, United Kingdom, 1995, Chap. 4,

p. 165.
9Spedding, G. R., and McArthur, J., “Span Efficiencies of Wings at Low Reynolds Numbers,” Journal of Aircraft, Vol. 47, 2010,

pp. 120–128. doi:10.2514/1.44247


10Spedding, G. R., and Hedenström, A., “PIV-based investigation of animal flight,” Experiments in Fluids, Vol. 49, 2009, pp.

749–763. doi:10.1007/s00348-008-0597-y
11Spedding, G. R., and Rignot, E.J.M., “Performance Analysis and application of Grid Interpolation Techniques for Fluid Flows,”

Experiments in Fluids, Vol. 15, 1993, pp. 417–430.


12Selig, M. S., Guglielmo, J. J., Broeren, A. P., and Giguere, P., Airfoils at Low Speeds, Soartech 8, Virginia Beach, 1989,

Chap.13, pp. 372.


13McCormick, B.W., Aerodynamics, Aeronautics, and Flight Mechanics, 2nd Ed., Wiley-India, United Kingdom, 1995, Chap. 4,

p. 115.

15
American Institute of Aeronautics and Astronautics

View publication stats

You might also like