You are on page 1of 19

Aero 301: Spring 2011 III.

5 3D Vortices & Biot–Savart Page 1

III.5 Vortices in 3D & The Biot–Savart Law


• We are finally ready to start thinking about the aerodynamics of 3D objects: wings that
do not extend to ±∞ into and out of the page. To do this we start by thinking about
what would happen if we were to impulsively start a motionless airfoil in a 2D world.
• We know 2 (seemingly) contradictory things
t < 0, No Motion
1. There will be vorticity/circulation associated with
Γ=0
the lift the moving airfoil generates but,
2. In the initial motionless state ω = 0 everywhere and
Dω /Dt = 0 so we should have ω = 0 everywhere for all time.
What gives?
t > 0, Airfoil moving right
• Dω /Dt doesn’t hold at the trailing edge (the
Kutta condition arises because of viscosity) so Γ>0 Γ=0
we do generate the vorticity and circulation Γ<0
required to produce the lift.
• But, besides that point, the flow remains inviscid
so no torque is applied to a large control volume
enclosing the wing and lots of space around it so the
circulation about that volume’s perimeter remains zero. bound vortices starting vortex

• So, for some region close to the wing to have positive circulation, some
negative circulation path must exist around a vortex not bound to the airfoil.
This other negative-Γ vortex is called a starting vortex. It has equal but
opposite strength to the net vortex strength that’s bound to the moving airfoil.
Aero 301: Spring 2011 III.5 3D Vortices & Biot–Savart Page 2

• Does this really happen? Yes!

Set aside this 2D picture for a few moments


and let’s move gingerly into 3D. . .
• Imagine that we make the simplest extension
from our 2D airfoil picture into 3D. This
would mean that the point vortices would
become lines that extend to ±∞ in
the y direction, into and out of the page.
• Next, imagine that instead of just vortex lines,
the 3D world can have vortex filaments, twisty
vortex strings that produce infinite vθ -type velocities
as distance from the filament goes to zero.
(You know these twisty vortices as tornados.)
• Working from the 3D incompressible Euler equations, we
could prove three vortex theorems developed by Helmholtz.
(But we won’t.)

H1 A vortex filament has constant strength, Γ along it’s length.


H2 A vortex filament cannot begin or end in a fluid. It must
end at a boundary, form a closed loop or extend to infinity.
H3 An inviscid fluid that is initially irrotational will remain
irrotational for all time.

• H1 and H2 do not apply in 2D


• H3 is nothing more than Dω
~ /Dt = 0.
Aero 301: Spring 2011 III.5 3D Vortices & Biot–Savart Page 3

• The ”closed loop” option of H2 explains the starting vortex that’s


observed in 2D flows. The two vorticies — one bound to the airfoil, the
other behind the airfoil — are just two bits of the same vortex loop.
• What about the rest of the loop?

As a finite wing (a 3D shape that


doesn’t extend to ±∞ in y) begins to
move, the Kutta condition generates a

n
vortex loop with one part of the vortex

pa
gs
bound to the wing, a starting vortex that

in
w
x

=
remains more or less at its starting r te

b
vo
position and two legs called wingtip ng
arti
vortices that connect the bound vortex St
to the starting vortex. Γ = constant
around loop
• We can observe the wingtip vortices and the starting
vortex. We know that the bound vortex exists
because we can measure the lift on the wing: L = ρ U∞Γ b
(although this ignores some nasty details). So, the
theoretical picture is in good agreement with observation.
• So, when a finite wing starts to move, it generates a starting vortex
and this remains pretty much where it originated (unless it’s near the
ground). As the wing moves, the bound vortex moves with it (so the
Kutta condition is satisfied) and the wingtip vortices get longer and
longer.
Aero 301: Spring 2011 III.5 3D Vortices & Biot–Savart Page 4

Top View Rear Views


Close to
trailing
edge

Side View Far from


trailing
edge
Aero 301: Spring 2011 III.6 Prandtl’s Lifting Line Page 5

• How do we analyze the effect of 3D vortex filaments?


• The 3D extension of vθ = −Γ /2π r is known as the

Γ d~s × r
Z
Biot–Savart Law: ~v = −
4π |~r|3

– d~s points along the vortex filament indicating the sense


of rotation (right-hand rule);
– ~r is the vector pointing from the point of interest (i.e.,
where the velocity induced by the vortex filament is
evaluated) to the point s along the vortex; and
– the integral is evaluated along the length of the vortex vortex
filament
which is usually ±∞ or a closed loop

• The velocity potential of this field satisfies Laplace’s


Equation in 3D as it must. ds

• When the Biot–Savart Law is applied to a straight vortex


filament that extends from ±∞, the velocity reduces to the r s
correct 2D behavior.
Use the setup to the right to verify this.
h
(Hints: r = h/ cos β and s = h tan β .)

β
• As another example, what is the velocity at the duθ
center of a circular vortex ring?
Aero 301: Spring 2011 III.6 Prandtl’s Lifting Line Page 6

III.6 Prandtl’s Lifting Line Theory, Downwash and Induced Drag


• What are the implications of H1, H2 and the Biot–Savart
Law on finite wings? How do we model finite wings?
Γ = constant
around loop
• Imagine that all the little γ ds vortices arrayed along the
camber line of a 2D airfoil are collected into a single Γ z
that’s placed at the quarter-chord point. y

n
x

pa
gs
• In 3D, this vortex cannot end so wingtip vortices

in
w
connect the bound vortex back to the starting

=
b
vortex that is still sitting on the runway, 1000
miles behind the wing.
w(y) < 0
• The bound vortex generates lift.
• The starting vortex is so far aft, it does not induce any
velocity near the wing.
• The wingtip vortices generate downwash: w(y) < 0
between themselves, including at the location of the
bound vortex.
• What is the downwash velocity induced by the pair of
wingtip vorices along the bound vortex? Calculate using
Biot–Savart. . .
Aero 301: Spring 2011 III.6 Prandtl’s Lifting Line Page 7

−Γ
w(y) =
π b [1 − (2y/b)2]

• The preceding expression is a bit scary. It implies we have infinite


downwash velocities at the wing tips. Even disregarding the
impossibility of infinite velocities, we expect our small-angle
approximations do not work and that the wingtips are probably
stalled.
What do we do about this?
• The next level of approximation allows Γ to vary
along the wingspan, despite the fact that Γ must
be constant along a vortex filament. If we let
Γ → 0 at the wing tips, maybe we can avoid infinite
downwash.
How do we have our cake and eat it too? Lifting Line
Γ1
• Because a single vortex filament must have Γ2
constant Γ , we simply stack a number of bound Γ3

vortices along the lifting line but allow them to


turn back into trailing vortices at different points
along the span. This allows for a varying Γ (y). Γ2
Γ3

Γ1
Aero 301: Spring 2011 III.6 Prandtl’s Lifting Line Page 8

• Why might Γ vary along the span? All the reasons that L′ can
vary for an airfoil: chord, angle of attack and zero-lift angle
can all vary along the span.

– Chord variations are called taper: c = c(y)


– Angle of attack variations are called twist: α = α (y)
– Zero-lift angle variations are called
aerodynamic twist: αL=0 = αL=0(y)
– And, as we will see, there is an induced angle of attack,
αi (y ) that decreases the effective angle of attack to less
than the geometrical angle of attack.

• Now we need to consider what multiple vortices mean for the


downwash velocity w(y) along the lifting line.

• Consider the diagram shown to the right that shows


the lifting line from directly upstream. The z
downwash at the point y0 due to each of the Γ1,A Γ2,A Γ3,A Γ3,B Γ2,B Γ1,B
half-infinite vortices is y
1 Γn y0
wn(y0 ) = − b/2 −b/2
2 2π (y0 − yn )
where n is an index for each of the half vortices and
yn is the y location of each of these. Each of the ‘A’
vortices is positive, each of the ‘B’ vortices is the
negative of the corresponding ‘A’.
Aero 301: Spring 2011 III.6 Prandtl’s Lifting Line Page 9

• In the picture the vortices 1A, 2A, 3B, 2B, and 1B induce
velocities down; 3A induces a velocity up. The different
directions is given by the signs of y0 − yn and the signs of Γn
• If we have a large number of very weak vortices we can do a
little calculus and say
dΓ (y) dΓ

1
dw(y0 , y) = − =− dy
4π (y0 − y) 4π (y0 − y) dy y

and, with this,


1 dΓ
Z b/2
1
w(y0 ) = − dy.
4π −b/2 y0 − y dy y

This is the net downwash at y due to all the vortices.


U∞
• The net effect of this downwash is to tilt the
incoming flow vector down the tilt angle is called w(y0)
an induced angle of attack, αi and, like w(y), this αi(y0)
angle depends on y.
• We define the induced angle of attack to be positive if
w < 0 (as it usually is) so
w(y0)
αi (y0 ) ≈ − if w(y) ≪ U∞
U∞
Aero 301: Spring 2011 III.6 Prandtl’s Lifting Line Page 10

• What are the effects of this tilt?

1. The effective angle of attack is less than the geometrical angle


of attack (the angle between the chord line and U∞ and this
leads to less lift at each station along the blade that you would
expect based on the geometrical angle of attack.
2. The aerodynamic force perpendicular to U∞ at any y0 is L′ cos(αi )
(i.e., it is decreased slightly because it is tilted back)
3. There is now an aerodynamic force in the direction of U∞
called induced drag: Di′ = L′ sin αi .

• To sort out the implications of all this, we need a way to determine


Γ (y). This function determines L′ and w at each station along the span
and needs to correctly reflect all the geometrical features of the wing.
• The strategy for finding Γ (y) is to equate two separate expressions for
L′ for each 2D airfoil section along the wing. First, the
Kutta–Joukowski Theorem gives at y0

L′ (y) = ρ U∞Γ (y0 ).

Second, the 2D airfoil characteristics give


1
L′ (y) = 2π [α (y0 ) − αi (y0 ) − αL=0(y0 )] × ρ U∞2 c(y0 )
2

• So, we have two different expressions for L′ at any y0. If we set these
equal to each other we can solve for the one thing we do not know for
a wing we have built, the Γ (y) distribution.
Aero 301: Spring 2011 III.7 Elliptical Wings Page 11

• Setting the two L′ espressions equal results in the


Fundamental Equation of Finite Wing Theory

" #
2 Γ (y0 ) 1 dΓ
Z b/2
1
= 2π α (y0 ) − αL=0(y0 ) − dy
U∞ c(y0 ) 4π U∞ −b/2 y0 − y dy y

• If we solve this equation for Γ (y0 ) we know the the lift at each
section and, from this, the lift on the wing.
• This whole process is very similar to the development of
thin-airfoil theory. Equating two expressions for L′ gives Γ (y0 )
and integrating Γ (y0 ) over the wingspan gives the overall lift.
• It is difficult to solve this equation (it is another integral
equation) so our solution procedure will again include a sine series
(this is nice because we would like to have Γ = 0 at the wingtips
to avoid infinite downwash.
• What is unfortunate about all this is that for a given wing at a
given U∞ more lift requires more Γ (via an increased geometrical
angle of attack). We see that increasing Γ also increases the
drag. However, increasing Γ also increases αi so, actually, Di
increases like L.
Aero 301: Spring 2011 III.7 Elliptical Wings Page 12

III.7 The Elliptical Lift Distribution

Solving the Fundamental Equation of Finite Wing Theory requires us to guess at a Γ (y) distribution and show
it’s a correct guess. (The same approach we used for the γ (x) distribution for thin airfoils.)

As a first guess we consider an elliptic distribution:

"  2 #1/2
2y
Γ (y) = Γ0 1 −
b

This distribution has circulation Γ0 at the root (y = 0) and Γ = 0 at the wingtips (which avoids the infinite
downwash problem).

First, let’s compute the downwash by taking the derivative dΓ /dy and performing the variable transformations:
2y/b = cos θ . With this we obtain
Γ0
w=−
2b
An elliptic Γ distribution produces uniform downwash.
Aero 301: Spring 2011 III.7 Elliptical Wings Page 13

• Is such a Γ distribution possible? Put it into the


fundamental equation of finite wing theory to verify. . .

• Yep! It works for an elliptic c(y) distribution if there is no


twist and no aerodynamic twist (i.e., α and αL=0 are
constants along the span):
"  2#1/2
2y
c(y) = c0 1 −
b

• We can integrate this chord distribution to find the


planform area, S and the aspect ratio, AR= b /S

π c0b 4b
S= and AR =
4 π c0

• We prefer to cite results in terms of b and AR rather than b


and c0 because not all wings are elliptical but all have an
unambiguious wingspan and a (nearly) unambiguous
planform area.
Aero 301: Spring 2011 III.7 Elliptical Wings Page 14

• Because we know Γ (y), then we can integrate across the span to


find the lift:

4U∞ b (α − αL=0)
Γ0 =
AR + 2

• And with w << U∞ such that cosαi ≈ 1 then αi can be written as:

πρ U∞2 b2(α − αL=0)


L=
AR + 2

AR
CL = 2π (α − αL=0)
AR + 2
Aero 301: Spring 2011 III.7 Elliptical Wings Page 15

• What about the induced drag? Integrating across the span gives

π
Di = ρΓ02
8

• The induced drag coefficient is

CL2
CDi =
π AR

The induced drag coefficient depends on the lift coefficient squared.


Aero 301: Spring 2011 III.8 Non-Elliptic Lift Distributions Page 16

III.8 Non-Elliptic Lift Distributions

In general, any combination of chord distribution, twist distribution, and aerodynamic twist distribution will
produce lift and induced drag.

To accommodate all the possible variations, a Fourier series approach is used that maintains Γ = 0 at y = ±b/2:

Γ = 2bU∞ ∑ An sin(nθ ) where cos θ = 2y/b
n=1

This series approach is nothing more than a generalization of the elliptic Γ distribution because the elliptic
distribution has (
Γ0 /2bU∞ n = 1
An =
0 n>1

Using this approach the odd coefficients, A1 , A3 , . . . , represent the symmetric variations in Γ that one typically
imagines for a “normal” wing.

The even coefficients represent asymmetric variations (i.e., when the left wing is different than the right
wing). This doesn’t seem to be very common at first but these are the terms that are used to model aileron
displacements that induce rolling moments.
Aero 301: Spring 2011 III.8 Non-Elliptic Lift Distributions Page 17

As before, we can think about the results we would get with a given set of An ’s without actually computing the
coefficients. The approximate lift and induced drag are given by integrations similar to those used for the
elliptic distribution.

Begin by integrating for the lift (watch for helpful orthogonality!)

CL = π A1 AR

The lift coefficient only depends on the first coefficient in the series. However, unlike thin airfoil theory, this
single-term result becomes a better approximation to the lift as the number of An’s computed is increased.
(See below.)

Similarly, for drag,


Aero 301: Spring 2011 III.8 Non-Elliptic Lift Distributions Page 18

N 2
CL2

An
CDi = (1 + δ ) where δ= ∑n
π AR n=2 A1

δ is always positive and, for a reasonably well designed wing, δ ≪ 1


so the preceding expression is often written

CL2
CDi = where e = (1 + δ )−1 ≈ 1 − δ
π e AR

The symbol e is selected because this number is an efficiency that is never greater than 100%.

Note that an elliptic lift distribution gives the minimum possible induced drag for a particular AR because it is
the only distribution that gives e = 1 because all of it’s An terms equal zero for n > 1.

To find the An ’s choose some number, N, of terms to keep in the sine series.
Aero 301: Spring 2011 III.8 Non-Elliptic Lift Distributions Page 19

Then, choose N points along the span and evaluate the fundamental equation at those N points using the
values of c, α , and αL=0 for each point. This gives N equations for the N unknown An’s.

You might also like