Anda di halaman 1dari 70

Finite Wing Theory

Wing Models
One may apply the results of 3-D potential theory in several ways. We first consider the
theory of finite wings.
We might start out by saying that each section of a finite wing behaves as described by
our 2-D analysis. If this were true then we would still find that the lift curve slope was 2
per radian that the drag was 0, and the distribution of lift would vary as the distribution of
chord. Unfortunately, things do not work this way. There are several reasons for this:

One explanation is that the high pressure on the lower surface of the wing and the low
pressure on the upper surface causes the air to leak around the tips, causing a reduction in
the pressure difference in the tip regions. In fact, the lift must go to zero at the tips
because of this effect. We will next see how and why we must model the 3-D wing
differently from 2-D.
If we were to take the naive view that the 2-D model
would work in 3-D, we might have the picture shown on
the right. If each section had the distribution of vorticity
along its chord that it had in 2-D, the lift would be
proportional to the chord, and would not drop off at the
tips as we know it must.
This sort of model does not conform to our physical
picture of what happens at the wing tips. And indeed, it does not satisfy the equations of
3-D fluid flow. The reason that this does not work is that in this case the streamlines are
not confined to a plane. They move in 3-D and the flow pattern is quite different.
We could go back to the governing equations and start simply with the linear Laplace
equation. By superimposing known solutions we could obtain a simple model of a 3-D
wing. We might start by superimposing vortices on the wing itself:

But this is no more than the strip theory model that did not work. The reason that this
model (which seems just to be a superposition of known solutions) is not adequate is that
it violates the governing equations in certain regions. The model does not satisfy the
Helmholtz laws since vorticity ends in the flow near the tips. Some additional
requirements must be imposed on the model. The requirements for such a model are just
the Helmholtz vortex theorems, discussed previously.
Our simple 3-D model above may be modified as shown below to satisfy the first of the
Helmholtz theorems.

In fact, as can be seen from the picture here, this vortex model is not too far from reality.

The downwash field and the existence of trailing vortices are not just some strange
mathematical result. They are necessary for the conservation of mass in a 3-D flow.

Air is pushed downward behind the wing, but this downward velocity does not persist far
from the wing. Instead it must move outward. The outward-moving air is then squeezed
upward outboard of the wing and the flow pattern shown above develops.
The trailing vortex is visualized by NASA engineers by flying an agricultural airplane
through a sheet of smoke.

The main effect of this vortex wake is to produce a downwash field on the wing.
This downwash field has several very significant effects:

It changes the effective angle of attack of the airfoil section. This changes the lift
curve slope and has many implications.

Induced drag: Lift acts normal to flow in 2D. This accounts for about 40% of the
fuel used in a commercial airplane, and as much as 80% of the drag in the critical
climb segments.

It produces interference effects that are important in the analysis of stability and
control.

The magnitude of the downwash can be estimated using the Biot-Savart law, discussed
previously.
When applied to our simple model with two discrete trailing vortices, the equation
predicts infinite downwash at the wing tips, a result that is clearly wrong. In fact, the
induced downwash is not even very large.
The failure of this simple model led Prandtl to develop a slightly more sophisticated one
in 1918. Rather than representing the wing with just one horseshoe-shaped vortex, the
wing is represented by several of them:

In this way the circulation on the wing can vary from the root to the tip. The strength of
the trailing vortex filaments is related to the circulation on the wing then by:
wake = wing
A vortex is shed from the wing whenever the circulation changes.
In the limit as the number of horseshoe vortices goes to infinity, the trailing wake is a
sheet of vorticity.

The trailing vortex strength per unit length in the y direction (vorticity) is the derivative
of the total circulation on the wing at that station. From this model, we can derive the
basic relations for finite wings.
The vorticity strength in the trailing vortex sheet is
given by:
= d /dy
and since the wing circulation changes most quickly
near the tips, the trailing vorticity is strongest in this
region. This is why we see tip vortices, and not a
complete vortex sheet, as in this NASA photo of an F111 in a 4-g turn. The vortices are visible in this picture
because the low pressure in this region lowers the
temperature and we see the condensed water vapor.

Lifting Line Theory

Basic Theory
We could try using 2-D flow results for each section, but correct them for the influence of
the trailing vortex wake and its downwash. This is the idea of lifting line theory.
We use the 2-D result that:
together with the relation:
to obtain:

But the angle of attack used here is reduced through the effects of downwash so that the
effective angle of attack is the true angle* minus the downwash angle:

Where the induced downwash, Wind, is given by the Biot-Savart Law:

Combining the expression for gamma:


with the expression for the downwash angle:

provides an integral equation for the circulation distribution along the wing.
Just as in thin airfoil theory, the integral equation can be solved by assuming a Fourier
series representation for the distribution.

Substitution of the expression for circulation into the integral equation leads to:

After integrating we have:

The solution of this equation for all values of y is not quite so easy as in the case of thin
airfoil theory where we could get closed form expressions for the An's. This is generally
done numerically. However, several interesting and simple results appear from this
analysis without ever actually computing the An's from the distribution of local angle of
attack. Some of these are discussed in the next section.

Elliptic Wing Results


If, for example, we represent the lift distribution with only a single term in the Fourier
series, then:

This represents an elliptic distribution of lift.


The downwash angle is, in this case:

The integral is constant when |y| < b/2.


In this domain:

Since the downwash distribution is constant the Cl distribution is just:

If the angle of attack is also constant along the wing (no twist) then the Cl is constant and
since:

Then in this case the section Clis equal to the wing CL and:

or:

Recall that this holds for unswept elliptical wings.

General Lift Distributions


If we are given the lift distribution we can compute the An's as we would with any Fourier
expansion. And once we know the Fourier coefficients, we may compute the downwash
distribution and the induced drag:

Substitution and evaluation of the definite integral** leads to:

This formula gives the downwash in the plane of the wing for arbitrary load distributions. For the
simple elliptical case, closed form solutions for the downwash and sidewash at the start of the
wake sheet exist. The simple relation for the velocity induced by an elliptic wing tailing vortex
sheet is:

Here, the variable Z is the complex coordinate y + iz and wo is the downwash at the wing root: y
= z = 0.
This formula permits computation of induced velocities behind a wing as they effect downstream
surfaces such as horizontal tails.

Note that the downwash is only constant in the plane of the wing and behind the wing. As we
move outboard of the wing or out of the plane of the wake, the downwash varies considerably
and there is a rather large upwash beyond the wing tips.
This downwash field produces several important effects. It changes the lift of surfaces in other
surfaces wakes. This is important in the analysis of airplane stability and the effectiveness of
horizontal tails. As can be seen from the downwash plot, the interference of a canard wake with a
wing is extreme: the wing lift is reduced behind the canard and the part of the wing outboard of
the canard has increased lift.
The downwash also produces induced drag as discussed in the next section.

Induced Drag and the Trefftz Plane


Fundamentals
The 2-D paradox that surfaces in inviscid flow produce no drag no longer applies in 3-D. The
downwash created by the trailing wake changes the direction of the force generated by each
section:

In three dimensions the force per unit length acting on a vortex filament is
Here, the local velocity V includes the component from the freestream and a component from the
induced downwash. This latter component produces a component of force in the direction of the
freestream: the induced drag.

The induced drag is related to the lift by:


From the results of lifting line theory for lift and downwash in terms of the Fourier coefficients of
the lift distribution:

so we have:

The induced drag is often written as:

This may be written in coefficient form as:

with the same definition of e. Note that e simply depends on the shape of the lift distribution. It is
called the span efficiency factor or Oswald's efficiency factor. Note also that the induced drag
force depends principally on the lift per unit span, L/b.
We can determine quickly, from the expression for induced drag above that drag is a minimum for
a given lift and span when all of the Fourier coefficients except the A1 term (which produces lift)
are zero. This corresponds to the elliptic loading case mention previously. In this case the
downwash is constant and e = 1.

Far Field Analysis and the Trefftz Plane


The analysis above works quite well for analyzing the drag of wings given the distribution of lift. It
was invented by Prandtl and Betz around 1920. It does involve a bit of hand-waving, though, and
requires some very approximate idealizations of the flow. For example: when we talk about the
downwash induced by the wake on the wing, just where on the wing do we mean? We assume a
single bound vortex line and compute velocities there, but a real wing does not have a single
bound vortex and the velocity induced by the wake varies along the chord. Fortunately, the
answers from this model are more general than the model itself appears. The induced drag
formulas can also be derived from very fundamental momentum considerations.

If the box is made large, contributions from certain sides vanish. In the limit as the box sides go to
infinity we obtain the following expressions for lift and drag:

Here, u, v, and w are the perturbation velocities induced by the wing and its wake. Note that the
drag only depends on the velocities induced in the "Trefftz Plane" -- a plane far behind the wing.
The drag can be expressed as the integral over the infinite plane of the perturbation velocities
squared. But, using Gauss' theorem it can be expressed as a line integral over the wake itself:

This simplifies the calculation of the drag.


The normalwash, Vn, is just the downwash if the wake is flat, but the downwash far behind the
wing, not at the wing itself.

Thus we would obtain the same expression as from lifting line theory if the downwash due to the
wake at the wing is half the downwash at infinity. This is indeed the case for unswept wings
modeled with a lifting line.

Nonplanar Wakes
All of the comments above apply to nonplanar wings as well as to simple planar ones. But we
must be careful about the assumed wake shape when evaluating forces in the far field. The
integrals above actually give the force on the wing and wake combination. Of course, in reality,
there is no force on the wake sheet, but if we assume a shape a priori, it is not likely to be a forcefree wake. However, since forces act in a direction perpendicular to the vortex, extending wakes
streamwise always yields a drag-free wake and nearly correct answers for drag using far field
methods.

Munk's Stagger Theorem

The result that the drag of a lifting system depends only on the distribution of circulation
shed into the wake leads to some very useful results in classical aerodynamics.
Perhaps the most useful of these is called Munk's stagger theorem. It states that:
The total induced drag of a system of lifting surfaces is not changed when the elements
are moved in the streamwise direction.
The theorem applies when the distribution of circulation on the surfaces is held constant
by adjusting the surface incidences as the longitudinal position is varied.
This implies that the drag of an elliptically-loaded swept wing is the same as that of an
unswept wing. It also is very useful in the study of canard airplanes for which the canard
downwash on the wing is quite complicated. Moving the canard very far behind the wing
does not change the drag, but makes its computation much easier. One may use the
stagger theorem to prove several other useful results. One of these is the mutual induced
drag theorem which states that: The interference drag caused by the downwash of one
wing on another is equal to that produced by the second wing on the first, when the
surfaces are unstaggered (at the same streamwise location).
These results are especially useful in analyzing multiple lifting surfaces.

Trefftz Plane Lift Derivation


We have discussed the calculation of drag based on the velocities induced in the Trefftz
plane, but can lift be calculated in a similar way?
The answer is not so easy. We start with the expression for force based on the momentum
equation.

Let's assume (naively, for now) that the contribution of each of these integrals goes to
zero on each side of the box, except for the back side, as the dimensions of the box are
increased. This leaves the contribution in the Trefftz plane due to the wake.
In the Trefftz plane, if we assume that all of the induced velocities are normal to the
plane, the lift becomes:

The evaluation of this integral seems straightforward.


But, it is not.
Consider the integral

when the wake is modeled simply by a pair of vortices.


The induced velocity, w, is given by:

Thus, the inner part of the above integral becomes:

So, the integral for lift is:

This looks exactly right; however, let's consider the same integral when the order of
integration is reversed:

The inner part of the above integral now becomes:

This integrand is antisymmetric as shown in the plot below. So, the integral for lift is: L =
0.

Now we have a paradox. We get two values for the same integral.
Actually, this is not a paradox; it is rather a function that is not Lebesgue integrable.
In order to evaluate an integral unambiguously, the function must satisfy two conditions:
1. It must be continuous, except at a countable number of points.
2. The integral of its absolute value must be finite.
To avoid this problem, the integral for lift can be first evaluated over finite limits. Taking
z from -A to A and y from -B to B we find:
L = (2 U / ) {(B+s) atan(A/(B+s)) - (B-s) atan(A/(B-s))
+ A/2 ln[ (A2+(B+s)2) / (A2 + (B-s)2)] }
The limit as A and B get large depends on the ratio of A to B.
When A>>B the value goes to 2 U s, when B>>A the value is 0 and when
A = B the value is U s. Thus the integral remains ambiguous when evaluated over an
infinite domain.
As noted by Larry Wigton of Boeing, this dilemma is resolved by using a different model
of the flow field. When the vertical velocity associated with thisvortex system is

integrated over the Trefftz plane, no ambiguities arise. But, the results are surprising.

The result is that the contributions from the finite length trailing vortices goes to zero.
The contribution from the bound vortex is found to be independent of the length of the
trailing vortices and is:

The starting vortex contribution is similarly independent of the trailing vortex length and
is equal to the bound vortex contribution. Thus, this lift is due to momentum flux, but not
from the trailing vortices.
We finally need to look at how the pressure term on the upper and lower sides of the
control volume is involved.
As might be expected, integrals once again are not unambiguous. They depend on the
relative sizes of the box sides, even though everything is infinite.
A careful analysis leads to the following basic results:
1. If the wake length is small compared with the box width and height then the lift is
associated with the momentum term of the starting and bound vortices.
2. If the wake length is large, and the box height is large compared with the width, then
the lift is associated with the momentum term of the trailing vortices.
3. If the wake is long and the width is large compared with the height, then the lift is
associated with the pressure terms on the top and bottom.

Nonplanar Wings
Even after the issues with infinite-domain integrals have been resolved, we must worry
about the assumed wake position. Although we could argue that streamwise wakes can
usually be used for far-field drag computataions, streamwise wakes can still support lift
forces. When the wing or wake is substantially nonplanar, these effects can be significant.

In fact, the vortex lift generated by highly swept wings can be estimated by far-field
methods only when the roll-up of the wake sheet is accurately computed. The alternatives
in such cases are to use near field methods or to compute the wake shape.

Computational Models
Panel Methods
Many computational models and analysis methods are based on linear three-dimensional
potential flow theory. These are discussed in the overview of panel methods in an earlier chapter.
In this section we take a look at the simplest panel method in more detail.

Weissinger Method
Weissinger theory or extended lifting line theory differs from lifting line theory in several
respects. It is really a simple panel method (a vortex lattice method with only one
chordwise panel), not a corrected strip theory method as is lifting line theory. This model
works for wings with sweep and converges to the correct solution in both the high and
low aspect ratio limits.
The version of this model used in the Wing Design program is actually a variant of
Weissinger's method: it uses discrete skewed horseshoe vortices as shown.

Each horseshoe vortex consists of a bound vortex leg and two trailing vortices. This
arrangement automatically satisfies the Helmholtz requirement that no vortex line ends in
the flow. (The trailing vortices extend to infinity behind the wing.)
The basic concept is to compute the strengths of each of the "bound" vortices required to
keep the flow tangent to the wing surface at a set of control points.

If the vortex of unit strength at station j produces a downwash velocity of AICij at station
i, then the linear system of equations representing the boundary conditions may be
written:

where {alpha} represents the angle of incidence of the sections along the span (assumed a
flat plates). If the section has camber, the local angle of attack is taken as the angle from
the zero lift line of the section.
The linear system of equations to be solved may also be written in terms of the angle of
attack at the wing root and the twist amplitude. For wings with a linear distribution of
twist (washout):

where:
{r} is a vector containing the root angle of attack as each element
{y} is the spanwise coordinate, varying from 0 at the root to b/2.
{} is the total twist (washout) in the wing from root to tip
Thus, the wing circulation distribution can be written as the sum of two distributions:
Since the section lift (lift per unit length along the span) is related to the circulation by:
The lift distribution can be expressed as:
where l1 and l2 are independent of the incidence angles and depend only on the planform
shape of the wing.
Since the lift coefficient of the wing, CL, is linearly related to the angle of attack we can
also write the lift distribution in the following form:

The first term is known as the additional lift distribution and the second term is called the
basic lift distribution. They scale linearly with the wing lift coefficient and the twist angle
respectively. Additional information on basic and additional lift distributions is available
in the section on wing design.

Wing Analysis Program


This Java application computes the lift and Cl distribution over a wing with sweep and
twist. To increase the angle of attack, click near the upper part of the plot; to reduce
alpha, click in the lower area.

Details:
The analysis is a discrete vortex Weissinger computation. Pitching moment is based on
the mean geometric chord and is measured about the root quarter chord point. The twist is
assumed linear and is taken to be positive for washout (tip incidence less than root
incidence).

Simple Sweep Theory


Lifting line theory works only for unswept wings.
Weissinger theory provides a means for computing the distribution of lift on swept wings,
but not the chordwise distribution of pressures.
Vortex lattice models, panel methods, and nonlinear CFD provide pressure distributions
on swept-back wings, but do not provide some of the insight that we can obtain with the
simpler models.
It is mostly for this reason, and partly for historical reasons that simple sweep theory is
interesting. It was invented by Buseman around 1935 and independently by R.T. Jones.
Consider an infinite wing as shown below.

If we ignore the effects of viscosity, and the wing is painted white so there is nothing to
distinguish one section from another, we can slide the wing sideways and we could not
tell that it was moving sideways. The air could not tell either, so the pressure distribution

would remain unchanged.

We have just created an infinite, obliquely-swept wing that is moving with respect to the
air at a speed:
We can use this fact to design a wing which can fly at a high speed with a pressure
distribution associated with a lower speed.
The main idea behind sweeping the wing is to reduce the effects of compressibility. The
component of the flow parallel to the wing is not effected by the presence of the wing; the
normal component is decoupled from the tangential component.

This is true not only according to linear flow theory, but also in the case of nonlinear
compressible flow with shock waves. It is an interesting exercise to show how the full
potential equations decompose into a normal term and a tangential term when one asserts
that nothing changes in the tangential direction. This idea is called simple sweep theory.
We can consider sections normal to the wing edges as operating in a flow with lower
Mach number and dynamic pressure. The effective normal Mach number is then:
but because of the reduced normal dynamic pressure, the section lift coefficient based on
this component of the freestream velocity must be increased if the total lift is fixed:

Furthermore, at a given angle of attack, the lift is reduced. The reduction of lift at a given
angle of attack for swept wings has important implications: the airplane incidence angle
must be higher, causing several problems for some aircraft on landing approach (e.g.
Concorde's drooped nose and long nose gear, F-8 variable incidence wing). The reduced
lift curve slope due to sweep can improve the ride quality in gusty air, however.
This basic idea permits subsonic sections to be used at supersonic freestream Mach
numbers or transonic airfoils to be used at Mach numbers higher than they would
otherwise be able to operate. This works quite well even up to Mach numbers well over
1.0. A recent airplane design that I worked on had a design Mach number of 1.4. The

airfoils were designed to operate at a Mach number of 0.7 (normal Mach number) with
the wing swept 60. Although the airplane lift coefficient was not meant to exceed 0.25,
the airfoils had a design Cl of 1.0 because of the reduced normal dynamic pressure from
simple sweep theory.
The reason for designing a supersonic wing with sweep and subsonic airfoil sections can
be seen in the results of thin airfoil theory which predicted no drag for subsonic sections,
but did indicate that supersonic airfoil sections would produce drag due to thickness,
camber, and lift.
Since the effect varies with cosine of the sweep angle, we expect that either forward or aft
swept wings would realize similar benefits. This is basically true, although, as discussed
in the section on wing design and forward swept wing, there are some important
differences. Similarly, wings with oblique sweep have been designed and tested. Further
discussions of oblique wing given in the section on supersonic wings

Forward-Swept Wings
Since sweep produces effects that vary with cos(sweep), we might expect that either
forward or aft sweep would yield the same results. To a first approximation, this is true;
but, many other considerations can be important in comparing designs with forward and
aft sweep. Historically these have led designers to adopt aft-swept wings for most
aircraft, but this was not universally true. The Hansa Jet was a forward-swept wing
business jet designed in the 1960's. Its forward swept wing permitted a larger cabin
without a wing spar interrupting the floor. Some sailplanes have slight forward sweep to
provide better visibility. Recently there has been renewed interest in the forward swept
wing concept for aerodynamic reasons and a demonstrator / research aircraft, the X-29
was built by Grumman for NASA, DARPA, and the Air Force.

Several aerodynamic advantages of the forward swept wing have been suggested. One of
the more interesting of these is illustrated below. The claim is that the lower surface of a
swept forward wing contributes a larger share of the total lift than the lower surface of an
aft-swept wing.

Although exaggerated in this figure, this effect is predicted and observed. It is due in part
to perturbation velocities induced by the 3-D thickness distribution and in part to the
velocities induced by streamwise vorticity.
Some of the advantages and disadvantages of forward sweep:
Advantages

Better off-design span loading (but with less taper: Cl advantage, weight penalty)
Aeroelastically enhanced maneuverability

Smaller basic lift distribution

Reduced leading edge sweep for given structural sweep

Increased trailing edge sweep for given structural sweep - lower CDc

Unobstructed cabin

Easy gear placement

Good for turboprop placement

Laminar flow advantages?

Disadvantages
Aeroelastic divergence or penalty to avoid it
Lower |Cl| (effective dihedral)

Lower Cn (yaw stability)

Bad for winglets

Stall location (more difficult)

Large Cm0 with flaps

Reduced pitch stability due to additional lift and fuse interference

Smaller tail length???

Oblique Wings
These results suggested to R.T. Jones that obliquely-swept wings would be the ideal shape for
supersonic aircraft wings. He first proposed the concept in the 1940's and flew flying wing models
at the first ICAS meeting in Madrid in 1958. A great deal of work has been done since on oblique
wing aircraft including design work by Boeing, General Dynamics, and Lockheed, wind tunnel
testing and analysis by NASA, and flight testing of models and piloted aircraft.
The picture below shows the AD-1 low speed oblique wing demonstrator.

Although one of the principal advantages involves reduced supersonic wave drag, the concept
has other merits.

When compared to an aircraft with symmetric variable sweep (such as the B-1 or F-111), the
oblique wing has little change in the aerodynamic center position. This keeps the stability at
reasonable levels and avoids large trim changes or complex fuel-pumping schemes.

In addition, several structural advantages have been suggested, especially for variable-sweep
aircraft: bending loads on the pivot are avoided, only a single pivot is required, the actuator loads
are reduced, and the straight carry-through structure reduces weight in other portions of the
aircraft.

An all-wing version of the oblique wing was first proposed by G. H. Lee in 1962. The
idea has been revived with the advent of active control systems and a recent artist concept
of the oblique flying wing is shown below.

Despite several questions about stability and control, this concept can be made to fly .

Oblique Wing Flight

This model weighed about 80 lb (36 kg) with a 20 ft (6 m) wing span and
was powered by two ducted fan engines. The aircraft is unstable and
controlled actively with a custom-built flight control computer. The project
was undertaken at Stanford University in 1991-1994, principally by Steve
Morris and Ben Tigner.

Aerodynamics of Slender Bodies


Sometimes we are not interested in generating lift, just reducing drag, and when we have
to reduce the drag of a given volume, the best shape is often a slender body -- and often
nearly a body of revolution. The aerodynamics of such shapes is quite different from
airfoils and wings, but follows some of the the same basic principles.
Bodies including fuselages are important because they produce drag, lift and moment.
They also produce important interference effects with wings and substantially change the
stability of an airplane.
The flow over more general fuselages and bodies can be predicted in much the same way
as the flow over airfoils and wings. Superposition of sources and doublets in the form of
panel methods or simpler forms (ala thin airfoil theory) are often used. Navier Stokes
equations are used when flow separation is suspected.

In the first part of this section, we examine the results of such calculations rather than the
methods themselves, while the second part briefly introduces the concept of slender body
theory.

Flow over Bodies


Some closed-form solutions for the potential flow over bodies of revolution are available
and are useful as reference results.
We noted that in 2-D the maximum velocity on an ellipse was given by:
umax/U = 1 + t/c
In 3-D the surface velocity over an ellipsoid of revolution is given by:

x is the distance from the midpoint in units so that the length is 2.0 and r is the radius (or
in these units, the ratio of diameter to length.)
The maximum velocity is given by:

The figure below (from Schlichting) illustrates the pressure distribution on bodies of
revolution. D/L = 0.1

Note that perturbation velocities are much smaller than in 2-D. These velocities may be
estimated by superimposing point sources. In this case for an ellipsoid:

Note that the maximum velocity is sensitive to the actual shape, with a paraboloid having
about 50% larger perturbations. The results from such a distribution of sources on the
axis, slightly underpredicts the velocity perturbations.
The figure below shows the pressure distribution on a typical fuselage shape with D/L =
0.09 computed by a source distribution on the x-axis. Note how the pressure falls near the
center of the cylindrical portion of the fuselage.

As indicated below, fuselages in inviscid flow produce a nose-up pitching moment when
the angle of attack is increased. This effect is destabilizing and is an important
consideration in the sizing of the horizontal tail.

Although the inviscid picture suggests that no lift is produced, the viscous flow actually
separates at the back of the fuselage, making the moment somewhat smaller and the lift
larger than predicted by inviscid theory. This lift produces induced drag and the fuselage
behaves as a low aspect ratio wing.
The figure below shows the effect of angle of attack on fuselage lift, drag, and moment
based on experimental data. Also shown is the estimated moment based on inviscid
theory.

Slender Body Theory


A simple theory can be used to estimate the aerodynamic characteristics of bodies that
vary slowly in X-direction: wings of very low aspect ratio and high sweep or slender

fuselages.

In such cases, the rate of change of all quantities in the x direction is small and the
governing equation becomes:

Note that the (1-M2) du/dx term has been dropped because we assume that everything
varies slowly in the X direction. The remaining equation is an equation for 2 dimensional
incompressible flow in the cross-plane, and since the Mach dependence has dropped out,
it is valid for both M<1 and M>1.

The 2-D cross flow can be computed using a conformal mapping method or even a 2-D
panel method. Note that in the cross flow plane, the flow is unsteady as the span of the
swept wing or the diameter of the fuselage changes as it cuts through the plane. So while
the solution to Laplace's equation still provides the correct velocity distribution, the
pressures must be computed using the unsteady Bernoulli equation.
One particularly simple and useful case is that of a highly-swept, low aspect ratio wing.

Using the slender body concept we can solve for the lift of a low aspect ratio wing.
Such low aspect ratio wings tend to produce constant downwash and thus nearly elliptic
loading until the angle of attack gets large.
The rate of change of momentum of the air in the cross-flow plane is equal to the force
per unit length on the wing:

If we consider a certain area in the cross plane


given a velocity w (
) then the force becomes:
It can be shown from 2-D unsteady theory (not discussed here) that the effective area , A

is given by a circle around the plate of diameter equal to the local span, Y.
Thus,

The total force on the wing is then given by:

This force acts normal to the wing plane. So for small angles of attack:

This shows that in the low aspect ratio limit, the lift curve slope differs from what lifting
line theory predicted by a factor of 2. An expression for lift curve slope derived from
second order corrections to lifting line theory is given by Jones:

where p is the ratio of wing semi-perimeter to wing span.


For a rectangular wing p = (b+c) / b = 1 + 1/AR so CL = 2AR / (AR+3).
This expression is in close agreement with experiment over a wide range of aspect ratios.
A similar analysis can be done for slender bodies of revolution.

This leads to the result that the lift produced by a body with a cut-off base is given by:

Note again that these results are independent of Mach number. The Prandtl-Glauert
correction still applies, but the reduction in forces due to effective stretching in the xdirection just cancels the increase in the velocities (1/1-M2) that would have been applied
in 2-D.

Wing Design

There are essentially two approaches to wing design. In the direct approach, one finds the
planform and twist that minimize some combination of structural weight, drag, and
CLmaxconstraints. The other approach involves selecting a desirable lift distribution and then
computing the twist, taper, and thickness distributions that are required to achieve this
distribution. The latter approach is generally used to obtain analytic solutions and insight into the
important aspects of the design problem, but is is difficult to incorporate certain constraints and
off-design considerations in this approach. The direct method, often combined with numerical
optimization is often used in the latter stages of wing design, with the starting point established
from simple (even analytic) results.
This chapter deals with some of the considerations involved in wing design, including the
selection of basic sizing parameters and more detailed design. The chapter begins with a general
discussion of the goals and trade-offs associated with wing design and the initial sizing problem,
illustrating the complexities associated with the selection of several basic parameters. Each
parameter affects drag and structural weight as well as stalling characteristics, fuel volume, offdesign performance, and many other important characteristics.
Wing lift distributions play a key role in wing design. The lift distribution is directly related to the
wing geometry and determines such wing performance characteristics as induced drag, structural
weight, and stalling characteristics. The determination of a reasonable lift and C l distribution,
combined with a way of relating the wing twist to this distribution provides a good starting point for
a wing design. Subsequent analysis of this baseline design will quickly show what might be
changed in the original design to avoid problems such as high induced drag or large variations in
Cl at off-design conditions.
A description of more detailed methods for modern wing design with examples is followed by a
brief discussion of nonplanar wings and winglets.

Wing Design Parameters

Span
Selecting the wing span is one of the most basic decisions to made in the design of a wing. The
span is sometimes constrained by contest rules, hangar size, or ground facilities but when it is not
we might decide to use the largest span consistent with structural dynamic constraints (flutter).
This would reduce the induced drag directly.
However, as the span is increased, the wing structural weight also increases and at some point
the weight increase offsets the induced drag savings. This point is rarely reached, though, for
several reasons.
1. The optimum is quite flat and one must stretch the span a great deal to reach the actual
optimum.
2. Concerns about wing bending as it affects stability and flutter mount as span is increased.
3. The cost of the wing itself increases as the structural weight increases. This must be
included so that we do not spend 10% more on the wing in order to save .001% in fuel
consumption.
4. The volume of the wing in which fuel can be stored is reduced.
5. It is more difficult to locate the main landing gear at the root of the wing.
6. The Reynolds number of wing sections is reduced, increasing parasite drag and reducing
maximum lift capability.

On the other hand, span sometimes has a much greater benefit than one might predict based on
an analysis of cruise drag. When an aircraft is constrained by a second segment climb
requirement, extra span may help a great deal as the induced drag can be 70-80% of the total
drag.
The selection of optimum wing span thus requires an analysis of much more than just cruise drag
and structural weight. Once a reasonable choice has been made on the basis of all of these
considerations, however, the sensitivities to changes in span can be assessed.

Area
The wing area, like the span, is chosen based on a wide variety of considerations including:
1. Cruise drag
2. Stalling speed / field length requirements
3. Wing structural weight
4. Fuel volume

These considerations often lead to a wing with the smallest area allowed by the constraints. But
this is not always true; sometimes the wing area must be increased to obtain a reasonable C L at
the selected cruise conditions.
Selecting cruise conditions is also an integral part of the wing design process. It should not be
dictated a priori because the wing design parameters will be strongly affected by the selection,
and an appropriate selection cannot be made without knowing some of these parameters. But the
wing designer does not have complete freedom to choose these, either. Cruise altitude affects the
fuselage structural design and the engine performance as well as the aircraft aerodynamics. The
best CLfor the wing is not the best for the aircraft as a whole. An example of this is seen by
considering a fixed CL, fixed Mach design. If we fly higher, the wing area must be increased by
the wing drag is nearly constant. The fuselage drag decreases, though; so we can minimize drag
by flying very high with very large wings. This is not feasible because of considerations such as
engine performance.

Sweep

Wing sweep is chosen almost exclusively for its desirable effect on transonic wave drag.
(Sometimes for other reasons such as a c.g. problem or to move winglets back for greater
directional stability.)
1. It permits higher cruise Mach number, or greater thickness or CL at a given Mach number
without drag divergence.
2. It increases the additional loading at the tip and causes spanwise boundary layer flow,
exacerbating the problem of tip stall and either reducing C Lmax or increasing the required
taper ratio for good stall.

3. It increases the structural weight - both because of the increased tip loading, and
because of the increased structural span.

4. It stabilizes the wing aeroelastically but is destabilizing to the airplane.


5. Too much sweep makes it difficult to accommodate the main gear in the wing.

See section 9.2.5 for more detail on simple sweep theory and the effect of sweep.
Much of the effect of sweep varies as the cosine of the sweep angle, making forward and aftswept wings similar. There are important differences, though as discussed further in the section
on forward swept wings.

Thickness
The distribution of thickness from wing root to tip is selected as follows:
1. We would like to make the t/c as large as possible to reduce wing weight (thereby
permitting larger span, for example).
2. Greater t/c tends to increase CLmaxup to a point, depending on the high lift system, but
gains above about 12% are small if there at all.
3. Greater t/c increases fuel volume and wing stiffness.
4. Increasing t/c increases drag slightly by increasing the velocities and the adversity of the
pressure gradients.
5. The main trouble with thick airfoils at high speeds is the transonic drag rise which limits
the speed and CL at which the airplane may fly efficiently.

Taper
The wing taper ratio (or in general, the planform shape) is determined from the following
considerations:
1. The planform shape should not give rise to an additional lift distribution that is so far from
elliptical that the required twist for low cruise drag results in large off-design penalties.
2. The chord distribution should be such that with the cruise lift distribution, the distribution
of lift coefficient is compatible with the section performance. Avoid high C l's which may
lead to buffet or drag rise or separation.
3. The chord distribution should produce an additional load distribution which is compatible
with the high lift system and desired stalling characteristics.
4. Lower taper ratios lead to lower wing weight.
5. Lower taper ratios result in increased fuel volume.
6. The tip chord should not be too small as Reynolds number effects cause reduced C l
capability.
7. Larger root chords more easily accommodate landing gear.

Here, again, a diverse set of considerations are important.


The major design goal is to keep the taper ratio as small as possible (to keep the wing weight
down) without excessive Cl variation or unacceptable stalling characteristics.
Since the lift distribution is nearly elliptical, the chord distribution should be nearly elliptical for
uniform Cl's. Reduced lift or t/c outboard would permit lower taper ratios.
Evaluating the stalling characteristics is not so easy. In the low speed configuration we must know
something about the high lift system: the flap type, span, and deflections. The flaps- retracted
stalling characteristics are also important, however (DC-10).

Twist
The wing twist distribution is perhaps the least controversial design parameter to be
selected. The twist must be chosen so that the cruise drag is not excessive. Extra washout
helps the stalling characteristics and improves the induced drag at higher CL's for wings
with additional load distributions too highly weighted at the tips.
Twist also changes the structural weight by modifying the moment distribution over the
wing.
Twist on swept-back wings also produces a positive pitching moment which has a small
effect on trimmed drag. The selection of wing twist is therefore accomplished by
examining the trades between cruise drag, drag in second segment climb, and the wing

structural weight. The selected washout is then just a bit higher to improve stall.

Wing Lift Distributions


As in the design of airfoil sections, it is easier to relate the wing geometry to its performance
through the intermediary of the lift distribution. Wing design often proceeds by selecting a
desirable wing lift distribution and then finding the geometry that achieves this distribution.
In this section, we describe the lift and lift coefficient distributions, and relate these to the wing
geometry and performance.

About Lift and Cl Distributions

The distribution of lift on the wing affects the wing performance in many ways. The
lift per unit length l(y) may be plotted from the wing root to the tip as shown below.

In this case the distribution is roughly elliptical. In general, the lift goes to zero at the
wing tip. The area under the curve is the total lift.
The section lift coefficient is related to the section lift by:

So that if we know the lift distribution and the planform shape, we can find the Cl
distribution.
The lift and lift coefficient distributions are directly related by the chord distribution.
Here are some examples:

The lift and Cldistributions can be divided into so-called basic and additional lift
distributions. This division allows one to examine the lift distributions at a couple of
angles of attack and to infer the lift distribution at all other angles. This is especially
useful in the process of wing design.
From the discussion of lifting line theory and Weissinger theory, we saw that the
distribution of lift could be written:

where:
ris the angle of attack at the root
is the twist angle
{l1} is the wing lift distribution with no twist and with r = 1
{l2} is the lift distribution at zero angle of attack and unit twist.
The lift distribution may also be written in a more conventional way:

Here, the distributions {la} and {lb} are the wing lift distributions with no twist at CL
= 1 and with unit twist at zero lift respectively. The first term, CL {la}, is called the
additional lift. It is the lift distribution that is added by increasing the total wing lift.
{lb} is called the basic lift distribution and is the lift distribution at zero lift.
Why is this useful? Consider the following example.

We can use the data at these two angles of attack to learn a great deal about the wing.
From the expression above:
or:
The additional lift distribution, CL {la} may be interpreted graphically as shown
below.

The additional lift coefficient distribution at CL = 1.0 is plotted below. Note that it
rises upward
toward the tip -- this is indicative of a wing with a very low taper ratio or a wing with
sweep-back.

The basic lift distribution is negative near the tip implying that the wing has washout.

Wing Geometry and Lift Distribution

The wing geometry affects the wing lift and Cl distributions in mostly intuitive ways.
Increasing the taper ratio (making the tip chords larger) produces more lift at the tips,
just as one might expect:

But because the section Cl is the lift divided by the local chord, taper has a very
different effect on the Cldistribution.

Changing the wing twist changes the lift and Cldistributions as well. Increasing the tip
incidence with respect to the root is called wash-in. Wings often have less incidence
at the tip than the root (wash-out) to reduce structural weight and improve stalling
characteristics.

Since changing the wing twist does not affect the chord distribution, the effect on lift
and Cl is similar.
Wing sweep produces a less intuitive change in the lift distribution of a wing.
Because the downwash velocity induced by the wing wake depends on the sweep, the
lift distribution is affected. The result is an increase in the lift near the tip of a sweptback wing and a decrease near the root (as compared with an unswept wing.

This effect can be quite large and causes problems for swept-back wings.
The greater tip lift increases structural loads and can lead to stalling problems.
The effect of increasing wing aspect ratio is to increase the lift at a given angle of
attack as we saw from the discussion of lifting line theory. But it also changes the
shape of the wing lift distribution by magnifying the effects of all other parameters.
Low aspect ratio wings have nearly elliptic distributions of lift for a wide range of
taper ratios and sweep angles. It takes a great deal of twist to change the distribution.

Very high aspect ratio wings are quite sensitive, however and it is quite easy to depart
from elliptic loading by picking a twist or taper ratio that is not quite right.

Note that many of these effects are similar and by combining the right twist and taper
and sweep, we can achieve desirable distributions of lift and lift coefficient.
For example: Although a swept back wing tends to have extra lift at the wing tips,
wash-out tends to lower the tip lift. Thus, a swept back wing with washout can have
the same lift distribution as an unswept wing without twist.
Lowering the taper ratio can also cancel the influence of sweep on the lift
distribution. However, then the Cldistribution is different.

Today, we can relate the wing geometry to the lift and Cldistributions very quickly by
means of rapid computational methods. Yet, this more intuitive understanding of the
impact of wing parameters on the distributions remains an important skill.

Lift Distributions and Performance


Wing design has several goals related to the wing performance and lift distribution. One
would like to have a distribution of Cl(y) that is relatively flat so that the airfoil sections
in one area are not "working too hard" while others are at low Cl. In such a case, the
airfoils with Clmuch higher than the average will likely develop shocks sooner or will
start stalling prematurely.
The induced drag depends solely on the lift distribution, so one would like to achieve a
nearly elliptical distribution of section lift. On the other hand structural weight is affected
by the lift distribution also so that the ideal shape depends on the relative importance of
induced drag and wing weight.

With taper, sweep, and twist to "play with", these goals can be easily achieved at a given
design point. The difficulty appears when the wing must perform well over a range of
conditions.
One of the more interesting tradeoffs that is often required in the design of a wing is that
between drag and structural weight. This may be done in several ways. Some problems
that have been solved include:
Minimum induced drag with given span -- Prandtl
Minimum induced drag with given root bending moment -- Jones, Lamar, and
others

Minimum induced drag with fixed wing weight and constant thickness -- Prandtl,
Jones

Minimum induced drag with given wing weight and specified thickness-to-chord
ratio --

Minimum total drag with given wing span and planform -- Kuhlman

... there are many problems of this sort left to solve and many approaches to the solution
of such problems. These include some closed-form analytic results, analytic results
together with iteration, and finally numerical optimization.
The best wing design will depend on the construction materials, the arrangement of the
high-lift devices, the flight conditions (CL, Re, M) and the relative importance of drag and
weight. All of this is just to say that it is difficult to design just a wing without designing
the entire airplane. If we were just given the job of minimizing cruise drag the wing
would have a very high aspect ratio. If we add a constraint on the wing's structural weight
based on a trade-off between cost and fuel savings then the problem is somewhat better
posed but we would still select a wing with very small taper ratio. High t/c and high
sweep are often suggested by studies that include only weight and drag.
The high lift characteristics of the design force the taper ratio and sweep to more usual
values and therefore must be a fundamental consideration at the early stages of wing
design. Unfortunately the estimation of CLmaxis one of the more difficult parts of the
preliminary design process. An example of this sensitivity is shown in the figure below.

The effect of a high lift constraint on optimal wing designs. Wing sweep, area, span, and
twist, chord, and t/c distributions were optimized for minimum drag with a structural
weight constraint.

Wing Design in More Detail


The determination of a reasonable lift and Cl distribution, combined with a
way of relating the wing twist to this distribution provides a good starting
point for a wing design. Subsequent analysis of this baseline design will
quickly show what might be changed in the original design to avoid problems
such as high induced drag or large variations in Clat off-design conditions.
Once the basic wing design parameters have been selected, more detailed
design is undertaken. This may involve some of the following:

Computation or selection of a desired span load distribution, then


inverse computation of required twist.

Selection of desired section Cpdistribution at several stations along the


span and inverse design of camber and/or thickness distribution.

All-at-once multivariable optimization of the wing for desired


performance.

Some examples of these approaches are illustrated below.

This figure illustrates inverse wing design using the DISC (direct iterative
surface curvature) method. The starting pressures are shown (top), followed
by the target (middle), and design (bottom); light yellow = low pressure and
green = high pressure. This is an inverse technique that has been used very
successfully with Navier-Stokes computations to design wings in transonic,
viscous flows.
Below is an example of wing design based on "fixing" a span load
distribution. When the 737 was re-engined with high bypass ratio turbofans, a
drag penalty was avoided by changing the effective wing twist distribution.

The details of the pressure distribution can then be used to modify the camber
shape or wing thickness for best performance. This sounds straightforward,
but it is often very difficult to accomplish this, especially when it takes hours
or days to examine the effect of the proposed change. This is why simple
methods with fast turnaround times are still used in the wing design process.

As computers become faster, it becomes more feasible to do full 3-D


optimization. One of the early efforts in applying optimization and nonlinear
CFD to wing design is described by Cosentino and Holst, 1986.
In this problem, a few spline points at several stations on the wing were
allowed to move and the optimizer tried to maximize L/D.

Although this was an inviscid code, the design variables were limited, and the

objective function simplistic, current research has included more realistic


objectives, more design degrees of freedom, and better analysis codes.

--but we are still a long way from having "wings designed by computer."

Nonplanar Wings and Winglets

One often begins the wing design problem by specifying a target C pdistribution and/or
span loading and then modifying the wing geometry (either manually, by direct inverse, or
by nonlinear optimization). In the case of planar wings, the elliptic loading results provide
a useful benchmark in the creation of target loadings. (For high aspect ratio wings, 2D
airfoil results may provide a useful reference for the chordwise loading.)
More complex methods for creating target Cp's are beyond the scope of this discussion,
but we have little guidance at all when the wing is nonplanar.
This section deals with the problem of optimal loading for nonplanar lifting surfaces. It is
easily generalized to multiple surfaces.
When the wing is not planar, many of the previous simple results are no longer valid.
Elliptic loading does not lead to minimum drag and the span efficiency can be greater
than 1.0.
Here we will describe a method for computing the minimum induced drag for planar and
nonplanar wings. First, consider the distribution of downwash for minimum drag. This can
be obtained by using the method of restricted variations as follows.

We consider an arbitrary variation in the circulation distribution represented by 1 and


2which do not change the lift:
L = U 1 + U 2 = 0.
This implies: 1 = - 2
If the drag was minimized by the initial distribution:
D = /2 w1 1 + /2 w2 2 = 0.
So, w1 = w2
That is, the downwash is constant behind a planar wing with minimum drag.

In the general case, with multiple surfaces or nonplanar wings, the same approach may
be used. In this case, the condition for constant lift is:
L = U 1 cos 1 + U 2 cos 2= 0.
where is the local dihedral angle of the lifting surface.
For minimum drag:
D = /2 Vn1 1 + /2 Vn2 2 = 0.
where Vnis the induced velocity in the Trefftz plane in a direction normal to the wake
sheet (the normalwash).
In this case, 1 cos 1 = - 2 cos 2
so, Vn = k cos .
The normalwash is proportional to the local dihedral angle. Thus, the sidewash on
optimally-loaded winglets is 0, for example.
We may then solve for the distribution of circulation that produces this distribution of
normalwash.

Alternatively, one may use a more direct optimization approach. With the circulation
distribution represented as the row vector, {} and the wake modeled as a collection of
line vortices of strength {w}, we may write the wake vorticity in terms of the surface
circulation, based on a discrete vortex model as shown below.

The drag is then given by: D = /2 {Vn} {}


where Vnis the normal wash in the Trefftz plane computed using the Biot-Savart law.
{Vn} is related to the circulation strengths by:
{Vn} = [VIC] {}
where [VIC] is a function of the geometry.
So, D = /2 [VIC] {} {}
The lift is also a function of the circulations:
L = U {} {cos }
with the local dihedral angle.
Finally, the optimal values of {} are given by setting
d ( D + (L-Lref) ) / di = 0 where is a Lagrange multiplier.
This problem is sometimes done as homework, but some results are summarized below:
When the wing/winglet combination is optimized for minimum drag at fixed span, it
achieves about the same drag as a planar wing with a span increased by about 45% of
the winglet height.
The wing lift distribution is as shown below with increased lift outboard compared with
the no winglet case.

This increased tip loading along with the extra bending moment of the winglet leads to
increased structural weight. When a bending moment constraint replaces the span
constraint, wings with winglets are seen to have about the same minimum drag as the
stretched-span planar wings. This is shown below.

Induced drag of wings with winglets and planar wings all with the same integrated
bending moment (related to structural weight). Note that solutions to the left of the span
ratio = 1.0 line are not meaningful.

The same approach may be taken for general nonplanar wake shapes. The figure below
summarizes some of these results, showing the maximum span efficiency for nonplanar
wings of various shapes with a height to span ration of 0.2.

Several points should be made about the preceding results.


1. The result that the sidewash on the winglet (in the Trefftz plane) is zero for minimum
induced drag means that the self-induced drag of the winglet just cancels the winglet
thrust associated with wing sidewash. Optimally-loaded winglets thus reduce induced
drag by lowering the average downwash on the wing, not by providing a thrust
component.
2. The results shown here deal with the inviscid flow over nonplanar wings. There is a
slight difference in optimal loading in the viscous case due to lift-dependent viscous drag.
Moreover, for planar wings, the ideal chord distribution is achieved with each section at

its maximum Cl/Cdand the inviscid optimal lift distribution. For nonplanar wings this is no
longer the case and the optimal chord and load distribution for minimum drag is a bit
more complex.
3. Other considerations of primary importance include:
Stability and control
Structures
Other pragmatic issues

Examples: MultiPlanes

Multiplanes include biplanes such as the Wright 1902 glider shown below.
Although the Wright brothers exploited the structural advantages of biplanes,
rather than the lower vortex drag for fixed span and lift, their motivation was
partly aerodynamic. Based on their own tests and those of Otto Lilienthal, it was
apparent that at very low Reynolds numbers (typical of test conditions used by
these pioneers) highly cambered, thin sections performed much better than thicker
sections, making the cable-braced Lilienthal designs or the Wright biplane
concepts especially attractive. Because of the low flight speeds required for
Lilienthal's take-offs and landings and for the power plants available to the
Wrights, the designs needed to be light and incorporate large wing areas. This
requirement was satisfied well with the biplane configuration.

The multiplane concept was taken to extremes by Phillips in 1904. The aircraft
shown below with 20 wings would have had a high span efficiency, but the very
low Reynolds number of each wing would lead to poor performance. The struts
and cables of early biplane designs also led to large parasite drag, so the effects of
improved span efficiency were not obvious. Several modern proposals for

cantilevered or semi-cantilevered biplanes have emphasized the lower vortex drag


of such configurations at the expense of structural efficiency, Reynolds number,
and fuel volume.
The induced drag of a multiplane may be lower than that of a monoplane of equal
span and total lift because the nonplanar system can influence a larger mass of air,
imparting to this air mass a lower average velocity change, and therefore less
energy and drag. For a biplane, if the two wings are separated vertically by a very
large distance, each wing carries half of the total lift, so the induced drag of each
wing is 1/4 that of the single wing. The inviscid drag of the system is then half
that of the monoplane.

In addition to the well-known advantages in vortex drag, the favorable


interference between two wings of a closely-coupled biplane can be used to
improve the section performance. The lower-than-freestream velocity at the
trailing edge of the forward wing and the new boundary layer on the downstream
wing can be exploited and some of the difficulties with lower Reynolds numbers
for the biplane as compared with a monoplane can be alleviated if not turned to
advantage. Gains in CLmax, width of laminar drag bucket, and drag divergence
Mach number at fixed t/c are possible with good multiple element section design.
As an example, a single fully-laminar section (100% laminar flow on upper and
lower surfaces) can support a CL of about 0.4. A 2-element wing can be designed
with an overall CL of about 0.75. This may help to explain the preference for
biplanes in the low Reynolds number world of insects.

Examples: Closed Systems

The aerodynamics of nonplanar wing systems that form closed loops are very
interesting. Such configurations include box-planes, ring wings, joined wings, and
"spiroid-tip" devices. Wings that form closed loops, such as the ring-wing
illustrated below, do not eliminate the "tip vortices" or trailing vortex wakes even
though the wing has no tips. Still, the vortex drag of the circular ring wing is just
50% that of a planar wing with the same span and total lift and the concept has
been studied at several organizations, including early aviation pioneers, a major
aircraft manufacturer, as well as several toy companies.

The Lockheed box-plane, shown below, achieves even greater drag reduction at a
given span and height than the circular ring wing (in fact the theoretical minimum

vortex drag) in a configuration with reasonable high-speed performance (note the


desirable transonic area-ruling) and some structural advantages. Fuel volume,
landing gear integration, CLmax penalties, and lower section Reynolds numbers are
some of the disadvantages for this concept.

Early attempts at ring-wing design were just as ambitious and successful. The
figure below is the Bleriot III.

The recently-patented "spiroid wing tip" produces a reduction in induced drag,


much like that of a winglet. However, its closed planform shape may make it
possible to reduce local lift coefficientsoften a problem for winglets.

Although a closed lifting system may eliminate the wing tips, it does not eliminate
the trailing vortex wake. In fact, the lift produced by the system can be directly
related to the velocities in the wake that lead to induced drag. These systems are

still interesting because one may add a constant circulation vortex ring to the
system without changing the wake. Such a constant strength vortex distribution
does not add any lift, but it may be used to produce moments without induced
drag penalties or to manipulate section lift coefficients in a desirable way.

Examples: Strut-Braced Wings

Aircraft concepts that employ auxiliary aerodynamic surfaces as struts to improve


both aerodynamic and structural efficiency have been studied extensively.
In joined-wing designs (below) the horizontal tail sweeps forward and joins the
main wing, forming a strut. The tail is then in compression, reducing wing
bending moments. If the tail is large enough to be positively loaded, some induced
drag savings is achieved, while if it is carrying a down-load, the closed loop
feature of the system minimizes trim drag.

Pfenninger's laminar designs with lifting struts exploit the nonplanar strut
geometry primarily for structural weight and stiffness, although some induced
drag reduction may be achieved.

Examples: Winglets and Tip Devices

The most common contemporary nonplanar wing configuration is the wing with
winglets, as seen below on the McDonnell-Douglas MD-11. These surfaces do
reduce induced drag for a given span, as well as providing a means of quickly
distinguishing the airplane from a DC-10. The MD-11 design includes small
downward winglets, while the 747-400 employs a full-chord single winglet, and
many other variations are possible.

A variant of the winglet concept, the C-wing is discussed later in this paper. It
involves adding a horizontal winglet extension (a wingletlet?) and has interesting
aerodynamic, structural, and control implications.

Examples: Nonplanar Wakes

The induced drag of a nonplanar system can be lower than that of a planar system of the
same lift and span. This is true even when the wing surfaces themselves are co-planar, but
their vortex wakes are not. Examples of this phenomena include:
America's Cup sailboat keels. Here the keel and rudder (or twin keel surfaces) are
coplanar, but due to the substantial leeway angle and longitudinal displacement of
the two surfaces, the wake downstream of the boat resembles that of a biplane
system and the induced drag is reduced substantially.
Crescent wings. This phenomenon was postulated as the reason for the distinctive
planform shape of some bird wings and fish fins, although the effect is almost
unmeasurable.

Split-Tips. This design was created to exploit the nonplanar wake geometry and is
discussed in more detail in a subsequent section of this paper.

Results: What is Possible?


Each of these configurations provides particular advantages and disadvantages, although
each benefits from some reduction in induced drag compared with the conventional
monoplane. The reduction in vortex drag is shown below for biplanes, boxplanes, and
winglets with varying ratios of height to span. These results were computed using an
optimizing vortex lattice code, but agree with classical solutions from Prandtl, von
Karman and Burgers, Cone, and Jones. Note that the boxplane achieves the lowest drag
for a given span and height, although winglets are quite similar. Considerable savings in
induced drag are possible for a fixed span if large vertical extents are permitted.

Of course, adding vertical surfaces such as winglets add wetted area and weight due to
higher bending moments, while the weight of a cantilevered biplane is increased since for
a fixed total area, the chords (and dimensional thickness) of each wing are halved. Jones
showed that with fixed integrated bending moment (a rough indicator of wing weight)
winglets produced about as much drag savings as planar tip extensions. More recent
analyses using more realistic weight estimation methods have yielded similar results (but
with much a less broad optimum).
For some applications, this discouraging result is not relevant since the aircraft must
operate with a span constraint, or because the structural arrangement is not simply
analyzed.

The figure below illustrates the effect of nonplanar wing shape on span efficiency. Each
of the geometries, shown in front view below, is permitted a vertical extent of 20% of the
wing span. Each design has the same projected span and total lift. The results were
generated by specifying the geometry of the trailing vortex wake and solving for the
circulation distribution with minimum drag. So, each of the designs is assumed to be
optimally twisted. This was done by discretizing the vortex wake and solving a linear
system of equations for minimum drag with a constraint on overall lift. Similar results for
a variety of shapes have been described by Cone, Munk, Letcher, Jones, and others.
The results illustrate the variability in span efficiency among these designs. Note the
relatively small gain for the diamond-shaped device and the wing with dihedral, while the
C-wing shape achieves essentially the same drag as the boxplane.

Nonplanar Wakes: The Split Tip


From among this list of possible designs, we choose two ideas to look at in a bit more
detail. The first concept is based on the notion that it is the shape of the wake, not the
shape of the wing that is important to the total vortex drag. By sweeping the trailing edge
of the wing sharply backward or forward and placing the wing at an angle of attack, one
may generate a wake shape that looks very much like the wake of a wing / winglet
combination. The difficulty here is that we must twist the wing or create a planform
shape that achieves the optimal load distribution that corresponds to this geometry.
Moreover, for reasonable wing planforms, the amount of out-of-plane wake deformation
is very limited. For this reason the potential gains associated with crescent-shaped wings
or wings with highly forward-swept trailing edges are very small (about 1% or less)
unless the wing has a very low aspect ratio.

To exaggerate this effect, a wing with the geometry shown below was created. The idea
here was to generate a shape whose potential span efficiency gain for a given amount of
out-of-plane deformation was large. Based on the previous figure, a split tip geometry for
the wake was selected as a shape that could be generated by wake deflection and the wing
planform shown below was investigated. The figure shows the planform shape and the
shape of the wake trace when the wing is at 9 degrees incidence. Based on this wake
shape, an induced drag savings of about 5% is possible when the wing is optimally
loaded, and more as the angle of attack is increased.

Of course, the wake does not trail from the wing in the streamwise direction and careful
computation of rolled-up wake geometry and inviscid drag shows that the effect of wakerollup is to roughly double the gain expected for a streamwise wake. The 11% increment
in span efficiency was significant and the concept was studied in more detail both
theoretically and experimentally. The figure below shows the computed wake geometry
and wing paneling used to compute vortex drag with the high-order panel code A502.

Two wings were constructed and tested at NASAs Ames Research Center. The first was
an untwisted planform with an elliptical chord distribution, unswept quarter chord line,
and an NACA 0012 airfoil section. The second wing of the same area and span, also
untwisted with a 0012 airfoil section, incorporated the split tip geometry. Both models
were designed to incorporate a sensitive internal balance so as to minimize support
interference. The figure below shows the ratio of lift to drag for each of these wings
confirming the predicted lower drag of the split tip geometry.

To further confirm the theoretical predictions, estimates of vortex drag and wake shape
were compared from calculations, balance data, and a detailed wake survey. From the
wake survey, an explicit estimate for the vortex drag can be obtained. This value agrees
well with the computed result and the balance data.
The results are intriguing, and although the configuration was selected to exaggerate a
particular effect rather than to serve as a good airplane wing, its application to aircraft,
propellers, and rotors is currently under investigation.

The C-Wing: A Novel Nonplanar Wing Configuration


From the survey of nonplanar wing geometries discussed previously, one is struck by the fact that
one need not produce a closed system such as the box plane to achieve essentially all of the
induced drag savings that this configuration offers. In particular, extending the upper part of the
box only 10% of the span inward from the tip achieves a span efficiency within about 1% of the
complete box. Thus, one could achieve the drag savings of the box plane without the Reynolds
number and fuel volume penalties of the two-wing design. Furthermore, the small horizontal tip
extensions have some interesting implications for airplane design. This section addresses the Cwing concept in a bit more detail and focusses on the application to a large commercial transport
design.

C-Wing Aerodynamics

The optimal loading of this lifting system is shown in the figure below. The
circulation of the main wing is carried onto the winglet so that the winglet is
loaded inward. When the horizontal extension is added to the winglet, forming the
"C" shape, the circulation is extended from the winglet as well, producing a
surface that is loaded downward for minimum induced drag at fixed total lift. It is
only when the lifting surface is extended to the centerline to form a box plane that
the upper wing can efficiently carry an upload. This is because, as mentioned
previously in connection with closed systems, we can superimpose a constant
circulation ring on the closed system to redistribute the lift without changing the
wake.
This download on the C-wing horizontal surfaces affects structural weight and
trim and the implications for aircraft configuration concepts was intriguing.

This configuration was independently "discovered" by a genetic algorithm that


was asked to find a wing of fixed lift, span, and height with minimum drag. The
system was allowed to build wings of many individual elements with arbitrary
dihedral and optimal twist distributions. The figure below depicts front views of
the population of candidate designs as the system evolves. On the right, the best
individual from a given generation is shown.

C-Wing: Application to Large Aircraft


The first application of this concept to an aircraft design study was in connection with recent
interest in very large civil transports. Many difficulties remain to be resolved for a successful large
conventional design. Airport and manufacturing constraints limit the span of a new large aircraft.
The location of the outboard engine is a problem, and the height of the vertical tail becomes
excessive. The C-Wing design was proposed as one way of addressing these issues and
perhaps improving the performance of such large aircraft. Additional sections of this paper
describe the evolution of the C-wing concept and include the following:

C-Wing Summary
The advantages of the C-wing configuration for a large capacity subsonic transport are
listed below. They include those directly associated with the nonplanar wing geometry
and those that arise indirectly from the overall configuration shown on previous pages.

Nonplanar Wing:
1. Reduced span or reduced vortex drag at fixed span
2. Efficient trim with short fuselage
3. Improved lateral handling (lower effective dihedral, reduced adverse yaw)
4. Potential for aeroelastic control: prevent aileron reversal, active flutter control
5. Reduced tendency for pitch-up, control at high alpha
6. Reduced vertical tail height
7. Possible reduction in wake vortex strength

Configuration:
1. Improved aero/structural performance through span loading, potential for
reduced wetted area
2. Effective use of redistributed wetted area reduces high lift system cost or TO
thrust / noise, potential for laminarization.
3. Some advantages of all-wing design with reduced risks: egress, windows,
growth, structure, acceptability
4. 2 wing-mounted engines reduce obstacle problem with outer engine / engine out
yaw
5. Single deck in wing facilitates loading, emergency egress

Disadvantages:
1. Details of emergency egress remain uncertain
2. Aerodynamics of thick inboard sections still an issue
3. Aeroelastics may be controllable but may need to be controlled

C-Wing Configuration Development


Using the C-Wing configuration, the span of an otherwise conventional large aircraft can
be reduced. Because the fuselage tends to be rather short on double deck configurations,
the horizontal tail location is not much farther aft than the wing tips making it possible to
consider using the C-wing as the primary pitch control surfaces. (The horizontal C-wing
surfaces provide more stability for a given area as they are not affected by the aft fuselage
flow field and are less affected by wing downwash. Moreover, they provide a positive
trimming moment when optimally loaded.) The removal of the horizontal tail makes the
use of aft-fuselage-mounted engines a possibility, eliminating some of the severe
problems with the original outboard engine location. Despite some attractive features,
however, the performance advantages for this configuration are not substantial, and
probably not worth the risk associated with the unconventional design.

As the number of passengers reaches 600-800, the possibility of including some


passenger cabin area inside the wing appears more attractive. For the C-Wing
configuration the wing span is reduced and the wing chord increased to maintain the
desired lifting area and structural support for the tip surfaces, making this idea even more
appealing. Furthermore, when a long empennage is no longer required for horizontal and
vertical tail surfaces, one is led to the rather unconventional large aircraft configuration
pictured below.

This design comprises a three-surface configuration providing a large allowable c.g.


range, with a relatively lightly loaded wing to simplify high-lift system requirements and
accommodate passenger cabins in the wing. The vertical and horizontal tip extensions
provide an efficient means of satisfying stability and control constraints.

As the design evolved to the tri-jet shown below, the wing span was increased, but
remained substantially lower than the conventional design. More efficient use was made
of the existing 777 fuselage area and the thick inner wing section was modified based on
an investigation of high-speed thick sections.
The basic idea in this conceptual design study was not to obtain the highest performanace
for this large aircraft, but rather to provide a feasible solution to the large aircraft
problem. The design addresses many of the problems that arise from the simple scalingup of the conventional design.

C-Wing Optimization and Initial Results

Studies at Stanford and Boeing were undertaken as part of NASAs Advanced Concepts
Program in 1995. Concurrently, initial sizing and optimization of the original concept
were pursued at NASAs Langley Research Center.
Each of these studies involved analysis and numerical optimization of the basic concept.
At Langley the FLOPS computer program was modified to handle this geometry. At
Stanford and Boeing, the PASS and ACSYNT codes were also modified to analyze this
design. Existing engine decks representative of modern high bypass ratio engines were
used rather than estimating the performance of future technology. This represents a rather
conservative approach. Additional analysis with ADP engines and laminar flow control
remains to be completed, but several aspects of the design suggest that gains from such
technologies may exceed those obtained with conventional designs.
The figure below shows the addition of C-Wing tips to the McDonnell-Douglas Blended
Wing Body concept. The addition of these surfaces would permit the BWB configuration
as currently envisioned to fly with positive static stability with no change to the
aerodynamic design of the highly-loaded, thick transonic wing. The added weight and
skin friction drag of these surfaces may be partly offset by a reduction in induced drag
and by the relaxed moment constraints on the main wing sections. Although the concept
remains to be studied in any detail, its implications for controllability and efficient trim of
this flying wing design are promising.

Anda mungkin juga menyukai