You are on page 1of 43

3D Aerodynamics and Finite Wings

AA200
Lecture 11
March 6, 2011

AA200 - Applied Aerodynamics

AA200 - Applied Aerodynamics

Lecture 11

Basic 3D Theory
We start our discussion of three dimensional aerodynamics by considering
the simple case of irrotational, inviscid flow. When compressibility can also
be neglected, we consider solutions to Laplaces equation in 3D:
2 = 0

Just as in the 2-D case, since this equation is linear, we might construct
solutions by superimposing known solutions. In 2-D we used sources and
vortices extensively to construct the flow over airfoils. In 3-D we also use
sources and vortices to model the flow over wings and bodies.
This section shows how this is done, starting with some results for some
fundamental 3-D singularities.
2

AA200 - Applied Aerodynamics

Lecture 11

Fundamental Singularities in 3D Potential Flow


One may derive fundamental solutions to Laplaces equation in 3-D, just
as we did in 2-D (although complex variables are not quite so useful).
3-D Source
It was easily discovered that the potential: = kr satisfied Laplaces
equation in 3-D. Since V = , the velocity associated with this solution
is directed radially with a magnitude: V = rk2 . It is easily shown that the
S
constant k is related to the volume flow rate, S by: k = 4
, so:
V =

S
4r2

. The velocity distribution associated with this 3-D source dies off as r2
rather than r as in the 2-D case.
3

AA200 - Applied Aerodynamics

Lecture 11

Point Doublet
Another basic solution, that has been used with some success in
supersonic aerodynamics programs is the point doublet, obtained by moving
a point source and sink together while keeping the product of their strength,
S, and separation, L, constant. With = SL, the velocity associated with
the point doublet is:
~ = cos r + sin
V
2r3
4r3
Vortex Filament
One of the most useful fundamental solutions to the 3-D Laplace
equation is that of a vortex filament. A vortex filament may be visualized as
a thin tube in which the flow has vorticity, . In the limit as the diameter
of the tube is made small, but the circulation, , is held fixed, this region of
vorticity is called a vortex filament. We often represent regions of vorticity
as vortex filaments or discrete vortex elements.
4

AA200 - Applied Aerodynamics

Lecture 11

Helmholtz Vortex Theorems


Helmholtz summarized some of the properties of vortex filaments, or
vortices, in 1858 with his vortex theorems. These three theorems govern
the behavior of inviscid three-dimensional vortices:
1. Vortex strength is constant
2. Vortices are forever (end on boundaries or form a closed path)
3. Vortices move with the flow
5

AA200 - Applied Aerodynamics

Lecture 11

Vortex strength is constant: A vortex line in a fluid has constant


circulation.

This can be proved by imagining a closed 3-D loop around the vortex
line as shown:

The integral around the closed loop from a to b to c to d to a cuts


through no vorticity so from Stokes theorem the integral is zero. But as
the slit is made very small the integral approaches the sum of the integral
from b to c and the integral from d to a. These are the local circulations
around the vortex line and so, the circulations must be constant along the
line.
Since the vortex strength is constant along the vortex line the strength
6

AA200 - Applied Aerodynamics

Lecture 11

cannot suddenly go to zero. Thus, a vortex cannot end in the fluid. It can
only end on a boundary or extend to infinity. Of course in an real, viscous
fluid, the vorticity is diffused through the action of viscosity and the width
of the vortex line can become large until it is hardly recognized as a vortex
line. A tornado is an interesting example. One end of the twister is on a
boundary; but at the other end, the vortex diffuses over a large area with
vorticity.

As discussed in the section on sources and vortices, singularities such


as vortices in the flow move along with the local flow velocity. Here,
interactions of the vortices in the trailing wake, cause them to curve around
each other and to form the nonplanar wake shown below.
7

AA200 - Applied Aerodynamics

Lecture 11

Image from Head 1982 in van Dyke, An Album of Fluid Motion, used
with permission.

AA200 - Applied Aerodynamics

Lecture 11

Biot-Savart Law
The Biot-Savart law relates the velocity induced by a vortex filament
to its strength and orientation. The expression, used frequently in
electromagnetic theory, can be derived from the basic equations for the
3D potential. The result is:

In the simple case of an infinite vortex we obtain the 2-D result:


||V || =

2r

In the case of a horseshoe vortex, the two trailing legs contribute:


9

AA200 - Applied Aerodynamics

Lecture 11

In the web notes you will find a simple subroutine to compute the
velocity components due to a vortex filament of length Gx, Gy, Gz with the
start of the vortex rx, ry, rz from the point of interest.

10

AA200 - Applied Aerodynamics

Lecture 11

Finite 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.
11

AA200 - Applied Aerodynamics

Lecture 11

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
12

AA200 - Applied Aerodynamics

Lecture 11

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
13

AA200 - Applied Aerodynamics

Lecture 11

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.

14

AA200 - Applied Aerodynamics

Lecture 11

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.
15

AA200 - Applied Aerodynamics

Lecture 11

The main effect of this vortex wake is to produce a downwash field on


the wing.
This downwash field has several very significant effects:
1. It changes the effective angle of attack of the airfoil section.
changes the lift curve slope and has many implications.

This

16

AA200 - Applied Aerodynamics

Lecture 11

2. 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
3. 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:
17

AA200 - Applied Aerodynamics

Lecture 11

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.
18

AA200 - Applied Aerodynamics

Lecture 11

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
19

AA200 - Applied Aerodynamics

Lecture 11

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 F-111 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.

20

AA200 - Applied Aerodynamics

Lecture 11

Lifting Line Theory


These are the basic concepts of finite wing theory. Mathematical
modelling of a lifting wing via the superposition of a number of horseshoe
vortices that obey Helmholzs laws yields a reasonable approximation to the
flow around a wing of high apect ratio. The key elements of this theory are
described as follows:
The main problem of finite-wing theory is the determination of the
distribution of airloads on a wing of given geometry flying at a given speed
and orientation in space. The analysis is based on the assumption that at
every point along the span the flow is essentially two-dimensional; thus the
resultant force per unit span at any point is that calculated for the airfoil
using two-dimensional thin airfoil theory, but at an angle of attack corrected
for the influence of the trailing vortex configuration that itself depends on
the spanwise variation of the lift distribution on the wing.
21

AA200 - Applied Aerodynamics

Lecture 11

Figure 1 presents a good summary of the various elements involved in


this theory that we have discussed so far. The key thing to understand is
that there is strong feedback between a lift distribution and its wake: the
spanwise derivative of the wing lift distribution determines the local strength
of the wake, which, in turn, determines the distribution of downwash at the
wing, which, again, determines the lift distribution along the span of the
wing. The key is to determine a distribution of lift/circulation along the
span of the wing which is in harmony with the downwash created by its
wake. Once this is accomplished all aerodynamic performance parameters
can be derived, including the total lift and drag of the wing, the downwash
distribution and the deviation of the produced lift distribution from the
optimum one. More importantly, lifting line theory can tell us about the
effect of variations in the wing planform and twist distributions on the
performance of the actual wing.

22

AA200 - Applied Aerodynamics

Lecture 11

Figure 1: Summary of Key Elements of Lifting Line Theory

Consider the unknown spanwise distribution of circulation to be given


by (y). The strength of the circulation at any point of the wake is then
23

AA200 - Applied Aerodynamics

Lecture 11

given by
d(y)
.
dy
The net effect of this trailing vortex system on any point on the bound
vortex is a downwash, w(y), whose magnitude at at each point is given by
the integrated effect of the continuous distribution of semi-infinite vorticity
over the range b/2 < y < b/2. Notice from the Figure that resultant
velocity at the wing has two components V and w, at each point, which
define the induced angle of attack
x(y) =

i(y) = tan1

w(y)
V

(1)

Using the Kutta-Joukwoski law, the force per unit span on the bound vortex
has magnitude V and it is normal to V , that is, it is inclined with respect
to the z-axis at an angle i. Using small angle approximations, these
24

AA200 - Applied Aerodynamics

Lecture 11

various quantities can be shown to be


w
V
L0(y) = V

i(y) =

Di0 (y) = L0i = w


We have shown that the net effect of the wake on the wing is the downwash
velocity which can be calculated as
Z 2b d(t)
1
w(y)
dt
i(y) =
=
P
dt
V
4V 2b y t

(2)

We had also seen that


cl(y) = cl ((y) L0(y) i(y))

(3)
25

AA200 - Applied Aerodynamics

Lecture 11

where cl is the lift curve slope (which could also be a function of y), (y)
is the geometric angle of attack, and L0 is the zero-lift angle of attack
(due to camber) of that particular cross-section. Since:
L0
cl = 1
2
2 V c(y)
where c(y) is the local chord, then
1
(y) = cl Vc( L0 i)
(4)
2
and plugging in Equation 4 into Equation 2 we obtain the governing equation
for (y) as
1
(y) = Vc cl
2

"

#
Z 2b 0
1
(t)dt
L0
P
4V 2b y t

(5)

26

AA200 - Applied Aerodynamics

Lecture 11

Elliptical Lift Distribution


Equation 5 is easily solved for c(y) if we were given (y). This procedure
involves the solution of an algebraic equation, rather than an integral
equation. The most important case of this type of solution procedure is
that of the elliptical lift distribution. Suppose that s represents the value of
the circulation at the plane of symmetry. Then, an elliptical lift distribution
can be written as
s
2

y
= s 1
.
(6)
b/2
The induced angle of attack created by this elliptical load distribution can
be computed by subtituting Equation 6 into Equation 2. One can then
show that
r
 2
Z 2b d 1 t
dt
b/2
s
i(y) =
P
dt
4V 2b
yt
27

AA200 - Applied Aerodynamics

Lecture 11

Using the Glauert trigonometric substitution that we have seen several times
by now, we can show that
Z
cos d
s
P
i(0) =
2bV 0 cos 0 cos
which can be shown to be equal to a constant
i =

s
2bV

(7)

Therefore, the angle of attack of every airfoil section on the wing is constant
(assuming no geometric twist). If we further assume that the sectional liftcurve slope is independent of y, both the sectional lift and drag coefficient
will also be independent of y. Therefore
cl = cl 0
28

AA200 - Applied Aerodynamics

Lecture 11

cdi

Di0
= cli
q c

Using these conditions, the product of cl 0c must vary elliptically, since


s
0

= Vs

cl 0c =

2s
V


1


1

y
b/2
2

2

1 2
= cl 0 Vc
2

y
b/2

Assuming that cl 0 is a constant (untwisted wing of identical crosssectional properties) this indicates that in order to obtain an elliptically
loaded wing, we must have an elliptical planform. If the planform were
not elliptical, an elliptical load distribution can still be achieved by making
sure that the wing be twisted in such a way that the product of cl 0c still
29

AA200 - Applied Aerodynamics

Lecture 11

varied elliptically. Notice that although this is fine in theory, extreme taper
ratios may lead to excessively high sectional lift coefficients in the outboard
wing sections that may exceed the local clmax and lead to stall.
The wing properties of this elliptical lift distribution can be computed
by integrating the spanload variations as follows

CL =

1
1
2
2 V S

b
2

2b

L0dy =

sb
2VS

The wing lift coefficient and the sectional lift coefficient are equal when
the sectional lift coefficients are constant along the span, and therefore,
CL = cl.
Solving for s in the expression above, and substituting into the
30

AA200 - Applied Aerodynamics

Lecture 11

expression for the induced angle of attack we find that


i =

cl
CL
=
AR
AR

and then, the wing induced drag coefficient is given by


CDi

CL2
= cdi = CLi =
AR

The expression for the lift curve slope is then given by


CL

cl
=
1 + cl /AR

31

AA200 - Applied Aerodynamics

Lecture 11

(y) Solution for Arbitrary Lift Distributions


Assume that you are given the geometry of the wing and the twodimensional lift curve slope of each of its cross sections. That is, the
quantities L0, c(y), and cl (y) are known. What is the best approach
to solve for (y)? The solution procedure should bring about a sense of
deja-vu from thin airfoil theory.
Let us transform the spanwise coordinate y into a polar variable
according to
t =
y

b
cos
2
b
cos 0
2
32

AA200 - Applied Aerodynamics

Lecture 11

Then,

X
1
An sin n0
(8)
(y) = g(0) = cls csV
2
n=1
where the constant outside the summation is included for convenience and
the s subscripts indicate values at the symmetry plane. Note that the An
coefficients are all non-dimensional, since the circulation has units of
L L/T . In addition
0

(t)dt =
=
and

d d
dt = g 0()d
d dt
X
1
cl csV
nAn cos nd
2 s
n=1

Z 2b 0
Z 0
X
(t)dt 1
cos nd
P
= cls csV
nAn P b
2
2b y t
2 (cos 0 cos )
n=1
33

AA200 - Applied Aerodynamics

Lecture 11

You may remember the value of the following principal value integral from
thin airfoil theory
Z
sin n0
cos n
=
P
sin 0
0 cos cos 0
which can be used to evaluate the integral and, substituting the result and
Equation 8 into Equation 5 we get
c cl X
c cl
sin no
( L0)
An sin n0 =
nAn
cscls
4b n=1
sin 0
n=1
X

(9)

which must be solved for the values of An. This is typically done with
the collocation method by truncating the series after a certain value of n
and evaluating the expression at n different values of 0 to create a linear
system of n equations in n unknowns An. Unfortunately, the only way to
check if your truncation was too crude is to repeat the procedure with a
higher value of n and compute the difference between the results.
34

AA200 - Applied Aerodynamics

Lecture 11

Assuming that we are able to solve the Equation 9 for the unknown
coefficients of the circulation distribution An, we can now use these values
to compute the lift and drag coefficients of the wing. Firstly, the sectional
lift and drag coefficients can be found to be
cl =

cdi

V cls cs X
=
An sin n
q c
c m=1

= cli =

c2ls c2s
4bc

(10)
!

An sin n

n=1

X
k=1

kAk

sin k
sin

which can be integrated to yield


Z
CL =

b
2

2b

clqcdy cls cs
=
q S
S

Z
0

b
An sin n sin d
2
n=1
35

AA200 - Applied Aerodynamics

Lecture 11

The integration is carried out easily by switching the orders of integration


and summation and noting that
Z


sin n sin kd =

for n 6= k
for n = k

Since k = 1 in all of the terms of the interal above, all terms vanish except
for the n = 1 term and therefore we have
cls csb
A1
CL =
4S

(11)

and therefore, the wing lift coefficient for an arbitrary symmetrical circulation
distribution is proportional to A1 and independent of all other Fourier
coefficients. For an elliptical lift distribution CL was proportional to s
which is the only surviving term in the Fourier expansion.
36

AA200 - Applied Aerodynamics

Lecture 11

Similarly, the coefficient of drag of the wing can be computed as follows


Z
CDi =

b
2

2b

c2ls c2s

cdi qcdy
=
q S
8S

kAnAk sin n sin kd

n=1 k=1

which can be shown to be


C Di =

c2ls c2s X

16S

nA2n

n=1

which proves that, for a given lift coefficient (A1 6= 0), the optimum lift
distribution (the one that yields minimum induced drag) is the one that
minimizes CDi . In other words, all Ai except for A1 must be equal to zero.
This corresponds to the elliptic lift distribution that we have been discussing
earlier.
37

AA200 - Applied Aerodynamics

Lecture 11

Interestingly, using the result for CL in Equation 11, one can show that
the induced drag coefficient for the wing is given by
CDi

 2

X
An
=
n
AR n=1
A1
CL2

which is simply the result for the elliptically loaded wing with a correction
factor that is sometimes called the span efficiency factor, e, defined as
 2

X
1
An
=
n
e n=1
A1
In the case of symmetric loadings, the situation is simplified considerably.
If we assume that (y) = (y), then all of the even coefficients of the
series expansions, A2, A4, ... are zero.
38

AA200 - Applied Aerodynamics

Lecture 11

Comparison with Experiment

Classical results on rectangular planforms reported by Prandtl can be


seen in the following Figures. Seven wings of aspect ratios between 1 and
7 were tested in a wind tunnel. Figure 2 shows the variation of CL vs the
geometric angle of attack. In Figure 2 the results have also been corrected
by the induced angle for AR = 5
39

AA200 - Applied Aerodynamics

Lecture 11

Figure 2: CL vs. and 0 for 1 AR 7

CL
0
=+

1
1

5 AR

40

AA200 - Applied Aerodynamics

Lecture 11

Figure 3 shows the drag polars for each of these wings, where

CD = CD0 + CDi

In Figure 3 these graphs have also been corrected to AR = 5 so the new


drag coefficient, CD0 becomes

CD 0

CL2
= CD +

1
1

5 AR

41

AA200 - Applied Aerodynamics

Lecture 11

0
Figure 3: CL vs. CD and CD
for 1 AR 7

The corrected Figures show that there is indeed significant departure


from the theory for aspect ratios under AR = 3. If AR is larger than
this value, significant departures from elliptical wing loading can still be
tolerated by this theory. In fact, it is typical to introduce the span efficiency
42

AA200 - Applied Aerodynamics

Lecture 11

factor, e, to account for some of these deviations so that


C Di

CL2
=
ARe

where e = 1.0 for a perfectly elliptically loaded wing, while it is smaller


(typically in the range 0.9 e 1.0) for wings with lift distributions that
deviate from the elliptical load.

43

You might also like