You are on page 1of 21

Swept Wings

Related terms:

Laminar Flows, Boundary Layer, Airfoils, Angle-of-Attack, Drag Coefficient, Lift


Coefficient, Vortex Sheet

View all Topics

Learn more about Swept Wings

Flow Control and Wing Design


E.L. Houghton, ... Daniel T. Valentine, in Aerodynamics for Engineering Students
(Seventh Edition), 2017

10.8 Reduction of Wave Drag


Aspects of wave drag reduction were covered in the discussion of swept wings in
Section 7.7 and of supercritical airfoils in Sections 8.1.1 and 10.2. In the latter
case it was found that keeping the pressure uniform over the upper wing surface
minimizes the shock strength, thereby reducing wave drag. A somewhat similar
principle holds for the whole wing-body combination of a transonic aircraft. This
was encapsulated in the area rule formulated in 1952 by Richard Whitcomb [145]
and his team at NACA Langley. It was known that as the wing-body configuration
passes through the speed of sound, the conventional straight fuselage, shown in
Fig. 10.37(a), experiences a sharp rise in wave drag. Whitcomb's team proved that
this rise in drag could be considerably reduced if the fuselage was waisted, as
shown in Fig. 10.37(b), in such a way as to keep the total cross-sectional area of
the wing-body combination as uniform as possible. Waisted fuselages of this type
became common in aircraft designed for transonic operation.
Figure 10.37. Application of the area rule for minimizing wave drag.

The area rule was first applied to a production aircraft in the Convair F-102A, the
USAF's first supersonic interceptor. Emergency application of the area rule became
necessary because of a serious problem revealed during the flight tests of the proto-
type aircraft, the YF-102. Its transonic drag was found to exceed the thrust produced
by the most powerful engine then available, which threatened to jeopardize the
entire program, considering that supersonic flight speed was an essential USAF
specification. The area rule was used to guide a major revised design of the fuselage.
This reduced the drag sufficiently for supersonic Mach numbers to be achieved.

> Read full chapter

Leading-edge Flow Separation


PAUL K. CHANG, in Separation of Flow, 1970

3 Bursting of the Leading-edge Vortices


The bursting of a vortex occurs in a swirling pipe as well as on a swept wing. The
phenomenon of the bursting of vortices is closely connected with flow separation.
The break-down of the burst into turbulence was once tentatively regarded as being
due to the separated boundary layer from the upper surface of the wing interfering
with the rolling-up of the vortex layer from the leading edge [11], but this factor
is no longer considered to be important to vortex bursting. The bursting of the
leading-edge vortices is briefly outlined in this section because of its practical
importance, and in order to enhance the understanding of flow separation. The
principal features of bursting of vortices have been outlined by Roy [12], Maskell
[13], and Jones [14]. Bursting refers to the structural change from a strong regular
spiral motion to a weaker turbulent motion, which can occur at some position along
a vortex [15].

The bursting occurs due to the positive pressure gradient bringing the low energy
fluid to rest, and the bursting can be eliminated by fluid suction applied just
downstream of the original position, as in the case of flow separation. The analogy
between a burst and a normal shock wave at supersonic speed is rather loose, but
both cases are associated with an instability in decelerating flow causing a sudden
increase of pressure.

Since the occurrence of the burst on a wing causes a loss of suction locally on the
flow surface and a modification of the position of separation of surface flow beneath
the vortex, the subject of bursting should be well understood for practical purposes.

3.1 Bursting of steady leading-edge vortices


A short outline of this subject is given in Incompressible Aerodynamics [16], edited
by B. Thwaites, and Jones [14].

In the past, investigations were made experimentally, using a water tunnel or a


low-speed wind tunnel. References to the experimental studies are: Spreiter and
Sacks [17], Werlé [18], Örnberg [19], Bartlett and Vidal [20], Fink and Taylor [21],
Lanbourne and Pusey [11], Elle [22], Marsden et al. [23], Hall [24], Garner and Walshe
[25], Werlé [26], Rogers et al [27], Squire [28], Jones [29], Hall [30], Garner and Cox
[31], Earnshaw [32], and Lawford and Beauchamp [33].

The theoretical investigations were made by Legendre [34], Adams [35], Brown and
Michael [36], Roy [37], and Mangier and Smith [39].

As the listed references show, much has been done experimentally and analytically,
but as yet the phenomena of bursting of the leading-edge vortices are not fully
understood. Qualitative observations were made in order to understand the flow
behavior and to provide a sound basis for analytical prediction of steady and un-
steady bursting. The water-tunnel observations were concerned with how sweeping
back the leading edge could transform the highly turbulent flow of the leading-edge
separated flow on a stalled unswept sharp-edged wing into the stable vortex-flow
characteristics of a highly swept wing. Furthermore, the structure of the vortex was
studied; velocity components, pressure and noise level were measured, and the
location of the vortex core was determined.

In the analysis, by approximating the free vortex sheets by two line vortices, the flow
is presented simply [34, 35, 36], but the analytical results were not in good agreement
with experimental data. Mangier and Smith's solution based upon the slender wing
theory display, numerically, most of the experimentally observed features.

3.2 Bursting of the leading-edge vortices on wings


The bursting of the leading-edge vortices on a delta-shaped wing takes place as
follows [15]. Consider a delta wing submerged in a subsonic fluid flow, as shown in
Fig. 6.

FIG. 6. Formation of laminar vortices for sharp-edged delta plate [15]

Each of the pair of vortices for a delta wing may be regarded as being generated by a
conically rolled vortex sheet from the leading edge, but the vortex can be considered
as a core of rotational flow fed by the vorticity from the leading edge, increasing
in size and strength with distance downstream. The vortex layers spring from the
leading edge S1 and form a line of secondary separation S2 on the upper surface.
The observations indicate that upstream of the burst, each stream line that enters
the vicinity of the vortex after bending around the leading edge has the form of a
nearly cylindrical spiral (Fig. 7).
FIG. 7. Axial filaments of dye flowing past sharp-edged delta plate. Water tunnel
velocity 2 in./sec [15]

Hall's [30] theory on vortex structure indicates that the rotational core consists of two
parts, i.e. an inviscid but rotational annulus, and an inner viscous core surrounded
by the annulus. The rotational annulus may be regarded as a region of approximately
constant swirl velocity, and the viscous core as the region in which the swirl velocity
falls to zero at the axis. The bursting appears to be a property of the core of the vortex,
and the axial flow within the core can be extremely sensitive to small retardations
of the external flow. The ratio of the swirling to the axial velocity component is a
very important factor for bursting. Nose-shape appears to have little effect on the
incidence at which vortex break-down occurs.

Measurements show that the total pressure at the axis is nearly constant along the
length of the vortex after a very steep fall close to the apex. The value of the total
pressure at the axis is very low due to the deficiency of rotational velocity caused by
viscous diffusion within the narrow inner core. The geometrical configuration of the
axis of the vortex on the wing is nearly straight and at an angle to the free stream
in the upstream portion of flow, but downstream, approaching the trailing edge, it
curves towards the free stream direction.

The various observations of burst at low velocity and Reynolds numbers indicate
that vortex flows from the sharp swept leading edge decelerate suddenly along the
vortex axis, deflect, and perform a regular whirling motion, then break-down of
turbulence occurs. The bursting is caused by an adverse pressure gradient along the
axis coupled with a low total pressure within the vortex core. On a wing, the axial
flow is easily brought to rest by pressure recovery associated with the trailing-edge
and the wake. In general, if the pressure recovery associated with the trailing edge
is not sufficient to stagnate the axial flow, a burst does not occur on the body
surface, although a burst may take place during a subsequent axial pressure rise due
to vorticity diffusion in the wake. Although the pressure rise in the outer regions of
flow surrounding a vortex is gradual, a sudden deceleration immediately ahead of
a burst appears to lie in the steeper pressure rise along a stream line near the axis.
Under certain conditions it is possible, depending upon the ratio of the rotational
to axial velocity components, to expand a vortex spontaneously and to provide the
pressure rise necessary for stagnation of the axial filament. The position of burst may
be determined by considering a balance of various independent factors dictated by
the geometry of the body. The burst location is sensitive to the pressure gradient
along the vortex. When the burst occurs upstream of the trailing edge, its position
depends upon a combination of incidence and leading-edge sweep back, but is
largely independent of Reynolds number. Once the conditions necessary to cause a
burst have been met, its final location may be determined by the extent to which fluid
from the turbulent region formed downstream of the burst can penetrate upstream
along the axis of the vortex. The observations reveal further, that for low values of
Reynolds number, between the positions of axial deceleration and the turbulent
break-down, a region of periodic flow exists, and the axial filament performs a
regular whirling motion. The vortex may expand in an axisymmetric manner about
the stagnation point in the axial flow, and because an axisymmetric arrangement
seems to be unstable, a strong tendency exists to collapse into a spiral configuration.
Therefore, the spiral configuration is a secondary feature of the process of bursting
[14, 15].

An investigation of the flow over an oscillating delta wing with 80° leading-edge
sweepback in a water tunnel [40] showed that considerable differences exist in the
position of vortices for steady compared to the unsteady motion, as qualitatively
shown in Fig. 8.

FIG. 8. Position of leading-edge vortex for a delta wing (pitching axis at 0·625 c0
from apex) [14]
With increase of incidence, the vortex gained its strength and moved inboard, but
did not reach the position of steady conditions until the maximum incidence was
attained. Conversely, with the decrease of incidence, the vortex lost its strength, but
moved inward beyond the position of steady flow.

> Read full chapter

Wing Theory
E.L. Houghton, ... Daniel T. Valentine, in Aerodynamics for Engineering Students
(Seventh Edition), 2017

7.7.2 Swept Wings of Finite Span


The yawed wing of infinite span gives an indication of the flow over part of a
swept wing, provided it has a reasonably high aspect ratio. But, as with unswept
wings, three-dimensional effects dominate near the wingtips. In addition, unlike
for straight wings, for swept wings three-dimensional effects predominate in the
mid-span region, which has highly significant consequences for the aerodynamic
characteristics of swept wings and can be demonstrated in the following way.
Suppose that the simple lifting-line model that was shown in Fig. 7.23 is adapted for
a swept wing merely by making a kink in the bound vortex at the mid-span position.
This approach is illustrated by the broken lines in Fig. 7.34. There is, however, a
crucial difference between straight and kinked bound vortex lines. For the former
there is no self-induced velocity or downwash whereas for the latter there is, as is
readily apparent from Eq. (7.1). Moreover, this self-induced downwash approaches
infinity near the kink at mid-span. Large induced velocities imply a significant loss
in lift.
Figure 7.34. Vortex sheet model for a swept wing. This uses the same physics as
in Prandtl's lifting-line theory, but the analytical effort required for this model is
substantially greater.

Nature does not tolerate infinite velocities, and a more realistic vortex-sheet model
is shown in Fig. 7.34 (full lines). It is evident from this figure that the assumptions
leading to Eq. (7.23) cannot be made in the mid-span region even for high aspect
ratios. Thus, for swept wings, simplified vortex-sheet models are inadmissible and
the complete expression in Eq. (7.22) must be used to evaluate the induced velocity.
The bound vortex lines must change direction and curve around smoothly in the
mid-span region. Some may even turn back into trailing vortices before reaching
mid-span. All this is likely to occur within about one chord from the mid-span.
Further away, conditions approximate those for an infinite-span yawed wing. In
effect, the flow in the mid-span region is more like that for a wing of low aspect
ratio. Accordingly, the generation of lift will be considerably impaired in that region.
This effect is evident in the comparison of pressure coefficient distributions over
straight and swept wings shown in Fig. 7.35. The reduction in peak pressure over
the mid-span region is shown to be very pronounced.
Figure 7.35. Comparison of pressure distributions over straight and swept-back
wings.

The pressure variation depicted in Fig. 7.35(b) has important consequences. First,
if it is borne in mind that suction pressure is plotted in the figure, it can be seen
that there is a pronounced positive pressure gradient outward along the wing. This
tends to promote flow in the direction of the wingtips, which is highly undesirable.
Second, since the pressure distributions near the wingtips are much peakier than
those further inboard, flow separation leading to wing stall tends to occur near the
wingtips first. For straight wings, on the other hand, the opposite situation prevails
and stall usually first occurs near the wing root—a much safer state of affairs because
rolling moments are smaller and flow over the ailerons, needed to control rolling
moments, remains largely attached. These difficulties make the design of swept
wings considerably more challenging than the design of straight wings.

> Read full chapter

The Anatomy of the Wing


Snorri Gudmundsson BScAE, MScAE, FAA DER(ret.), in General Aviation Aircraft
Design, 2014

Method 2: Empirical Estimation for Swept Wings


Raymer [5] also presents the following statistical expression to estimate the Oswald
efficiency of swept wings. It has limitations similar to Equation (9-89):

(9-90)

Brandt et al. [4] present the following expression to estimate the factor:

(9-91)

where

LE = leading edge sweep angle


t max = sweep angle of the maximum wing thickness line

> Read full chapter

Companion Computer Programs


TUNCER CEBECI, in Analysis of Turbulent Flows, 2004

10.6 Differential Method with CS Model: Infinite Swept-Wing


Flows
This program, called BLP2ISW, is also the extension of BLP2 to the calculation of
infinite swept-wing equations for incompressible laminar and turbulent flows as
discussed in Problem 8.6. Again two subroutines are added to BLP2 and changes
are made to the eddy viscosity subroutines. Subroutine COEF2 includes the coef-
ficients of the z-momentum equation and subroutine SOLV2 is the same solution
algorithm used in BLP2H and given in Section 10.10.

For three-dimensional turbulent flows, the eddy viscosity formulas require changes
to those for two-dimensional flows. Here they are defined according to Eqs. (5.7.4)
and (5.7.5).

Sample calculations for an infinite swept wing having the NACA 0012 airfoil cross
section with a sweep angle of = 30°, an angle of attack of = 2°, chord Reynolds
number Rc = 5 × 106 and transition location at x/c = 0.10 are presented in the
accompanying CD-ROM. See also Problem 8.7.

> Read full chapter

Companion Computer Programs


Tuncer Cebeci, in Analysis of Turbulent Flows with Computer Programs (Third
Edition), 2013

10.6 Differential Method with CS Model: Infinite Swept-Wing


Flows
This program, called BLP2ISW, is also the extension of BLP2 to the calculation of
infinite swept-wing equations for incompressible laminar and turbulent flows as
discussed in Problem 8.6. Again two subroutines are added to BLP2 and changes
are made to the eddy viscosity subroutines. Subroutine COEF2 includes the coef-
ficients of the z-momentum equation and subroutine SOLV2 is the same solution
algorithm used in BLP2H see subsection 10.13.1.

For three-dimensional turbulent flows, the eddy viscosity formulas require changes
to those for two-dimensional flows. Here they are defined according to Eqs. (5.7.4)
and (5.7.5).
Sample calculations for an infinite swept wing having the NACA 0012 airfoil cross
section with a sweep angle of  = 30°, an angle of attack of  = 2°, chord Reynolds
number Rc = 5 × 106 and transition location at x/c = 0.10 are presented on the
companion site, store.elsevier.com/components/9780080983356. See also Problem
8.7.

> Read full chapter

Algebraic Turbulence Models


Tuncer Cebeci, in Analysis of Turbulent Flows with Computer Programs (Third
Edition), 2013

5.7.1 Infinite Swept Wing Flows


The accuracy of the CS turbulence model of this section and other models has been
investigated for several infinite swept wing flows, as discussed in [71]. Here we
present a sample of results taken from those studies and discuss first the results
for the data of Bradshaw and Terrell [72] and then for the data of Cumpsty and
Head [73].

Data of Bradshaw and Terrell


This experiment was set up especially to test the outer-layer assumptions made in
extending the boundary-layer calculation method of Bradshaw et al. [74] from two
dimensions to three [75]. Measurements were made only on the flat rear of the
wing in a region of nominally zero pressure gradient and decaying cross flow. See
the sketch in Fig. 5.24a. Spanwise and chordwise components of mean velocity and
shear stress, and all three components of turbulence intensity, were measured at a
number of distances x = 0, 4, 10, 16 and 20 in. from the start of the flat portion
of the wing (Fig. 5.24). The surface shear stress, measured with a Preston tube, was
constant along a generator to the start of the flat part of the wing, except for a few
inches at each end and except for small undulations of small spanwise wavelength
caused by residual nonuniformities in the tunnel flow.
Fig. 5.24. Results for the relaxing flow of Bradshaw and Terrell: (a) wall cross-flow
angle and local skin friction, (b) velocity profiles, (c) cross-flow angle distributions.
The symbols denote the experimental data, the solid line the numerical solutions of
Cebeci [76] and the dashed line the numerical solutions of Bradshaw et al.

[74]

Figure 5.24 shows the calculated results (solid lines) with experimental results (sym-
bols) and those obtained by Bradshaw’s method [74] (dashed lines). The cross-flow
angle which represents the departure of the velocity vector within the bound-
ary-layer from the freestream velocity vector was computed from

(5.7.7a)

The above formula becomes indeterminate at y = 0; however, with the use of


L’Hopital’s rule, it can be written as

(5.7.7b)

The streamwise component of the local skin-friction coefficient was calculated


from

(5.7.8)

with and Us given by

(5.7.9a)

(5.7.9b)
Here and denote the wall shear values in the x- and z-directions, respectively,
obtained from the solution of the infinite swept wing equations.

Data of Cumpsty and Head


In this experiment [73] the boundary-layer development was measured on the rear
of a wing swept at 61.1°. The boundary-layer separated at about 80% chord. The
measured profiles were affected by traverse gear “blockage,” probably because of
upstream influence of disturbance caused to the separated flow by the wake of the
traverse gear.

Figure 5.25 shows a comparison of calculated and measured streamwise velocity


profiles us/Us and streamwise momentum thickness 11 defined by

Fig. 5.25. Comparison of calculated (solid lines) and experimental (symbols) results


for the data of Cumpsty and Head on the rear of a swept infinite wing.

(5.7.10)

where us/Us is calculated from

(5.7.11)

The results in Fig. 5.25 show good agreement with experiment at two x-stations.
However, with increasing distance they begin to deviate from experimental values
and at x = 0.650 ft, the agreement becomes poor.

The above results indicate what was already observed and discussed in relation to
the shortcomings of the Cebeci-Smith algebraic eddy-viscosity formulation, that is,
it requires improvements for strong adverse pressure gradient flows. As discussed
in subsection 5.4.2, the improvements to this formulation were made for two-di-
mensional flows by allowing in the outer eddy-viscosity formula to vary. A similar
improvement is needed to the formulation for three-dimensional flows.
> Read full chapter

Algebraic Turbulence Models


TUNCER CEBECI, in Analysis of Turbulent Flows, 2004

5.7.1 Infinite Swept Wing Flows


The accuracy of the CS turbulence model of this section and other models has
been investigated for several infinite swept wing flows, as discussed in [71]. Here
we present a sample of results taken from those studies and discuss first the results
for the data of Bradshaw and Terrell [72] and then for the data of Cumpsty and Head
[73].

Data of Bradshaw and Terrell


This experiment was set up especially to test the outer-layer assumptions made in
extending the boundary-layer calculation method of Bradshaw et al. [74] from two
dimensions to three [75]. Measurements were made only on the flat rear of the
wing in a region of nominally zero pressure gradient and decaying cross flow. See
the sketch in Fig. 5.24a. Spanwise and chordwise components of mean velocity and
shear stress, and all three components of turbulence intensity, were measured at a
number of distances x = 0, 4, 10, 16 and 20 in. from the start of the flat portion
of the wing (Fig. 5.24). The surface shear stress, measured with a Preston tube, was
constant along a generator to the start of the flat part of the wing, except for a few
inches at each end and except for small undulations of small spanwise wavelength
caused by residual nonuniformities in the tunnel flow.
Fig. 5.24. Results for the relaxing flow of Bradshaw and Terrell: (a) wall cross-flow
angle and local skin friction, (b) velocity profiles, (c) cross-flow angle distributions.
The symbols denote the experimental data, the solid line the numerical solutions of
Cebeci [76] and the dashed line the numerical solutions of Bradshaw et al. [74].

Figure 5.24 shows the calculated results (solid lines) with experimental results (sym-
bols) and those obtained by Bradshaw's method [74] (dashed lines). The cross-flow
angle which represents the departure of the velocity vector within the bound-
ary-layer from the freestream velocity vector was computed from

(5.7.7a)

The above formula becomes indeterminate at y = 0; however, with the use of


L'Hopital's rule, it can be written as

(5.7.7b)

The streamwise component of the local skin-friction coefficient cfs was calculated
from

(5.7.8)

with ws and Us given by

(5.7.9a)

(5.7.9b)

Here wx and wz denote the wall shear values in the x- and z-directions, respec-
tively, obtained from the solution of the infinite swept wing equations.
Data of Cumpsty and Head
In this experiment [73] the boundary-layer development was measured on the rear
of a wing swept at 61.1°. The boundary-layer separated at about 80% chord. The
measured profiles were affected by traverse gear “blockage,” probably because of
upstream influence of disturbance caused to the separated flow by the wake of the
traverse gear.

Figure 5.25 shows a comparison of calculated and measured streamwise velocity


profiles us/Us and streamwise momentum thickness 11 defined by

Fig. 5.25. Comparison of calculated (solid lines) and experimental (symbols) results
for the data of Cumpsty and Head on the rear of a swept infinite wing.

(5.7.10)

where us/Us is calculated from

(5.7.11)

The results in Fig. 5.25 show good agreement with experiment at two x-stations.
However, with increasing distance they begin to deviate from experimental values
and at x = 0.650 ft, the agreement becomes poor.

The above results indicate what was already observed and discussed in relation to
the shortcomings of the Cebeci-Smith algebraic eddy-viscosity formulation, that is,
it requires improvements for strong adverse pressure gradient flows. As discussed
in subsection 5.4.2, the improvements to this formulation were made for two-di-
mensional flows by allowing in the outer eddy-viscosity formula to vary. A similar
improvement is needed to the formulation for three-dimensional flows.

Full Three-Dimensional Flows


To illustrate the evaluation of the CS model for full three-dimensional flows, we con-
sider two flows corresponding to an external flow formed by placing an obstruction
in a thick two-dimensional boundary-layer (data of East and Hoxey) and an external
flow on a prolate spheroid at an incidence angle of 10° (data of Meier and Kreplin).

Data of East and Hoxey


Figure 5.26 shows a schematic drawing of East and Hoxey's test setup in which a
wing is placed in a thick two-dimensional boundary layer [77]. The strong pressure
gradients exposed by the obstruction caused the boundary layer to become three-di-
mensional and to separate. The measurements were made in the three-dimensional
boundary layer upstream of and including the three-dimensional separation.

Fig. 5.26. Schematic drawing of East and Hoxey's test setup.

Figure 5.27 shows a comparison between calculated and measured velocity profiles
on the line of symmetry (Fig. 5.27a) and off the line of symmetry as described in [76].
In general the agreement with experiment is satisfactory.
Fig. 5.27. Comparison of calculated (solid lines) and measured (symbols) velocity
profiles (a) on the line of symmetry and (b, c, d) off the line of symmetry for the East
and Hoxey flow.

Data of Meier and Kreplin


Meier and Kreplin's data correspond to laminar, transitional and turbulent flow on
a prolate spheroid with a thickness ratio of 6 for a Reynolds number of 6.6 × 106
[78–80]. As discussed in [81], the calculations for this flow were made for freestream
velocities of 45 and 55 m/s corresponding to natural and imposed transition. To
account for the transitional region between a laminar and turbulent flow, the
right-hand sides of Eqs. (5.7.4) are multiplied by the intermittency factor tr, defined
by Eqs. (5.3.18) and (5.3.19). Since detailed and corresponding correlation formulas
for three-dimensional transitional flows are lacking, the same expression was used
for three-dimensional flows by using the local similarity assumption with ue in Eqs.
(5.3.18) and (5.3.19) replaced by the total velocity.

The experimental data of Meier et al. consists of surface shear stress magnitude and
direction vectors and velocity profiles over a range of angles of attack. Figure 5.28a
shows a comparison of calculated surface shear stress vectors in laminar flow at =
10°. The magnitude of the shear stress vector is proportional to the shear intensity.
The agreement between the calculation and measurements on the windward side
is generally good, although there are some differences that are partly due to the
use of inviscid potential flow in the calculations, whereas the measured pressure
distribution shows viscous-inviscid interaction effects. It is clear that the laminar flow
is separated on the leeward side of the body at some distance aft of the nose. The
origin or nature of the high shear intensities leeward of the separation line cannot
be determined from calculations because calculations based on external flow that is
purely inviscid is not expected to account for strong interactions.

Fig. 5.28. (a) Measured (→) and calculated (→) distributions of wall shear stress vectors
(cf = w/1/2 u 2) for laminar flow and (b) for laminar, transitional and turbulent flow
on a prolate spheroid at = 10° [81].
Figure 5.28b shows wall shear vectors for laminar, transitional and turbulent flow
with natural transition. In general, the calculated and measured results are in
agreement with discrepancies (which are small) confined to the region close to
the specified transition. More quantitative comparison with the imposed transition
experiment is afforded by Fig. 5.29 which displays circumferential distributions of
wall shear stress at four axial locations. The calculated results display the correct
trends and are within 15% of the measured values with discrepancies tending to
diminish with downstream distance. A sample of the velocity profiles is shown in
Fig. 5.30 and corresponds to x/2a of 0.48 and 0.73, and again the agreement is
within or very close to the error bounds of the measurements, except in the regions
where the inviscid velocity distribution differed from the measured one. Additional
comparisons of calculated and experimental data are given in [81].

Fig. 5.29. Measured (dashed line) and calculated (solid line) resultant wall shear stress
values on the prolate spheroid at = 10° [81].

Fig. 5.30. Comparison of calculated (solid lines) and measured (symbols) streamwise
us/Us and crossflow us/Us velocity profiles [81].
> Read full chapter

Mining the ‘far side’ of technology


to develop revolutionary aircraft proto-
types: the Defense Advanced Research
Projects Agency (DARPA) approach
J.R. Wilson, in Innovation in Aeronautics, 2012

11.5.2 1980s
• Stealth Fighter,

• Tacit Blue (Whale) Stealth Bomber testbed, new stealth approach leading to
B-2 Bomber,
• Joint STARS,

• X-29 Forward Swept Wing Aircraft Technology,

• Pilot's Associate,

• Materials Technologies for the F-15 and F-16,

• Low Probability of Intercept Airborne Radar,

• Teal Rain high-altitude, long-endurance, extended-range ISR/Target Acquisi-


tion UAV,
• Condor long-range UAV, all-composite honeycomb structure, autonomous
controls, high-altitude aerodynamics, fuel-efficient propulsion system –
considered Global Hawk conceptual prototype,
• X-wing Rotor Systems Research Aircraft,

• No Tail Rotor (NOTAR) helicopter.

> Read full chapter

Wing problems
T.H.G. Megson, in Aircraft Structures for Engineering Students (Sixth Edition), 2017

28.2.3 Swept wing divergence


In the calculation of divergence speeds of straight wings, the flexural axis was taken
to be nearly perpendicular to the aircraft's plane of symmetry. The bending of such
wings has no influence on divergence, this being entirely dependent on the twisting
of the wing about its flexural axis. This is no longer the case for a swept wing, where
the spanwise axes are inclined to the aircraft's plane of symmetry. Let us consider
the swept wing of Fig. 28.4. The wing lift distribution causes the wing to bend in
an upward direction. Points A and B on a line perpendicular to the reference axis
deflect by approximately the same amount, but this is greater than the deflection of
A , which means that bending reduces the streamwise incidence of the wing. The
corresponding negative increment of lift opposes the elastic twist, thereby reducing
the possibility of wing divergence. In fact, the divergence speed of swept wings is
so high that it poses no problems for the designer. Diederich and Budiansky, in
1948, showed that wings with moderate or large sweepback cannot diverge. The
opposite of course is true for swept-forward wings, where bending deflections have
a destabilizing effect and divergence speeds are extremely low. The determination
of lift distributions and divergence speeds for swept-forward wings is presented in
Bisplinghoff, Ashley, and Halfman.3

Figure 28.4. Effect of Wing Sweep on Wing Divergence Speed

> Read full chapter

ScienceDirect is Elsevier’s leading information solution for researchers.


Copyright © 2018 Elsevier B.V. or its licensors or contributors. ScienceDirect ® is a registered trademark of Elsevier B.V. Terms and conditions apply.

You might also like