Sei sulla pagina 1di 19

THE FINITE WING

9.1. The Finite Wing. TTp to this point we have considered only twotlimensional flow and have carried that development up to its current
stabus as a coherent phase of aerodynamics. Now we shall turn our
atkention to the finite wing and see what effect freeing the tips has on its
performance.
The basic mechanism of flight was understood and practiced (with
gliders) a long time ago. Many theorists became interested in its problems, and a decade before the first powered flight Lanchester (born 1868)
in England postulated the type of flow that would be experienced by a
finite wing. I n a paper in 1894 which later led to the publication of his
book "Aerodynamics"* in 1907 (Ref. 9.1) he stated that the high pressure beneath the wings would spill out around the tips into the lowpressure region above the wings, forming vortices which would stream
out behind the wing. These vortices would (he continued) roll up into
two main trailing vortices of opposite sign, located one behind each tip,
and would be deflected do\vnward. Further, their effect upon the flow
a t the wing would tip the resultant force vector back, causing a component of the lift t o become induced drag. It is evident that Lanchester
clearly understood this phenomenon when it is noted that in 1897 he
secured a patent covering the use of end plates a t the wing tips to minimize the spillage there-six years before the Wright brothers' flight!
Although Lanchest,er's work was mathematical in scope, his presentation was not, and it remained to Pra,ndt,l (born 1875) t o extend the work
of Lan~hest~er
into the Prandtl lifting-line theory (Ref. 9.2) presented in
191 1 and developed in tlhe following pages.
9.2. The Trailing Vortices. As Lanchester pointed out, the important
difference in the flow pattern about two- and three-dimensional wings is
traceable t o the difference in spanwise lift distribution, which is in turn
traceable to the disposition of the circulation. The two-dimensional
wing has constant circulation along the span. When we represent the
airfoil by a vortex, its effect ahead of the wing tending to increase the
angle of attack is exactly balanced by the effect behind the wing tending
to decrease it.
* Especially recommended to the student for its historical value.
181

182

B A S I C W I N G AA-D AIRFOIL THEORY

On the other hand, the three-dimensional wing has a diminution of


circulation toward the wing tip until, when the tip is reached, the circulation is zero. At each change of circula.tion a vortex is shed, and the
shed vortices, being of similar sign, roll up into a single trailing vortex
somewhat inside each wing tip. The net result is that the vortex pattern
behind the wing is no longer similar to that ahead of the wing. The

from bound

2%

(b)

FIG.9.1. Downwash velocity for infinite ( a ) and finite (b) wings. (The finite-wing vdue of
2wo is for a. first-approximation elliptic wing.) (Reproduced by permission from "Principles
of Aerodynamics" by James H . Dwinnell, published by McGraw-Hill Book Company, Inc..
1949.)

vertical induced velocities ahead and behind the wing no longer cancel,
as the air behind the wing is affected by the downthrust on the wing
circulation and the trailing vortices, while the air ahead of the wing is
affected almost entirely by the upward component of the \iring circulation
(see Fig. 9.1). (The effect of the trailing vortices is very small ahead of
the wing.) Finally, the whole field behind the wing has a downward
inclination whose vertical component is called the downwash, and at the

T H E FIlVITE W I N G

183

wing enough inclination results materially t o tip the resultant force


backward and create a drag.
A great contribution to the theory for a complete wing was Prandtl's
statement that each section of the wing acts as though it is an isolated
two-dimensional section a t an angle of
attack a,,. This condition, known as
Prandtl's hypothesis, is in general responsible for the existence of a simple
theory for a complete wing of finite
span. There are, however, certain
7
I
j
conditions when the spanwise flow
becomes so pronounced that it is not
applicable, as, for instance, when
severe lateral pressure gradients arise
with a swept-back wing. Since these
conditions become of interest in determining the stalling characteristics of FIG.9.2. Growth of shed vortices into
tip vortex, showing cancellation of comwings, they offer an irritating obstacle ponents
final reduced span.
to practical design a t the current
time. They are due, of course, to the fact that curved flow ensues,
owing to the above reasons, and the actual effective airfoil sections are
undefined.
Another great contribution to the mathematical representation of this
field was Prandtl's statement that the trailing vortices could be considered as part of one vortex which
extends across the wing span and
is closed a long way behind the
wing by a starting vortex (Fig. 9.3
and Sect. 4.2). This complete
vortex representation will enable
us t o calculate the induced velocity
a t any point near a wing as soon as
FIG. 9.3. Representation of starting "orwe develop relations to parallel in
tex and a change of lift.
three dimensions the simple relation of Eq. (4.3).
But first a few words about this three-dimensional vortex, which now
has a finite length to the axis of rotation. Our previous vortex was flat
(two-dimensional) and its core a disk. The concept of three dimensions
requires a vortex of finite length and, of course, a finite-length core as well.
Helmholtz (1821-1894) in 1858 formulated vortex laws that will assist
us in studying the three-dimensional vortex as well as the vortex pattern
of a finite wing. Briefly these were:

184

BASIC JVIA-G AIVB A I R F O I L T H E O R ) .

1. Vortex filaments either form closed curves or extend t o the fluid


boundaries.
2. The circulation remains constant throughout the length of a vortex.
3. The particles of fluid composing a vortex remain with that vortex
indefinitely.
Sir William Thomson (Lord Kelvin, 1824-1907) is usually credited
with the additional law, "Circulation remains constant with time."
Let us see how well actual conditions meet these la\vs.
Flow pictures made (Ref. 9.4) a t the instant air begins t o move over a
wing a t some lifting angle of attack show that a vortex, parallel t o the
wing span, leaves the wing trailing edge and proceeds downstream. This
vortex is called the starting vortex and is of strength equal t o that of the
lifting, or bound, vortex. The sum of the t ~ v ois thus zero, satisfying
the Thomson law.
The starting and bound vortices may be very simply demonstrated
with a razor blade and a basin of water. The razor blade, held half
immersed so that it represents a wing a t some lifting angle, is moved
briskly through the water. A starting vortex immediately leaves the
blade. As long as the motion is constant, no additional vortices appear,
but any increase of angle produces a second starting vortex. When the
motion is stopped, the bound vortex comes off the "wing." The starting
vortex, by connecting the ends of the two trailing vortices of the wing
vortex system, enables that system to meet the first Helmholtz law.
The system meets the second law by actually having constant circulation
in each vortex section. The third law is not quite satisfied, since the
system is continually picking up air and thrusting it into the trailing
vortex system. Indeed, in view of the molecular action taking place in
any gas there is doubtless a considerable exchange of molecules between
those in the core and in free air.
After the starting vortex has moved downstl-eam to a point where its
influence near the wing is negligible, there remaifis only the bound vortex
with two trailing vortices in the region. The shape of this combination
resembles a horseshoe, and ordinarily the system is so called, instead of
identifying its components.
Since the lift vector of a circular cylinder goes through its axis, the
horseshoe vortex system is most accurate if the bound vortex is assumed
t o run through the wing center-of-pressure line.
We see directly that our simple horseshoe vortex field, having its
vortex strength determined from L = pVI'b, fits in well with the actual
conditions. Silverstein, Katzoff, and Bullivant (Ref. 11.l) examined
this phenomenon by comparing the flow field about a Clark Y as determined by the method of Chap. 8 and the vortex lams of Chap. 4. It was

185

THE FINITE WING

discovered that a t a distance of one chord from the wing trailing edge the
downwash angles were matched within 0.3"by a simple point vortex field.
While we have just discussed the case of constant circulation across
the span, i t may be seen by reference to Fig. 9.4 that any symmetrical
(or, indeed, unsymmetrjcal) spanwise distribution of lift may be simulated by a number of superimposed closed vortex systems. In actuality,
these small vortices, shed a t every change of circulation on the span,
roll up into the tip vortices and move them slight,ly toward the plane of

---,
/

- - --

Actual lift
distribution

(a)

FIG.9.4. Theoretical vortex pattern for arbitrary distribution of circulation along the span.
roll-up of shed vortices neglected. (Reproduced hy permission from "Principles of Aerodynamics" by James H. Dwinnell, published by McGraw-Hill Book Company, Inc., 1949.)

symmetry. In addition the whole field aft of the wing is displaced


vertically downward by the bound vortex.
The manner in which the circulation varies along the span is of great
interest to the aerodynamicist since it will be shown that one particular
variation (elliptic) results in minimum induced drag, others producing
from 2 to 15 per cent more. All variations of the circulation are amenable to mathematical treatment, but for simplicity we shall consider four:
elliptic (produced by an untwisted wing with elliptic planform; uniform
(not actually produced by any planform, but of such simple mathematical
scope that i t helps in understanding the other cases); rectangular-wing
type; and tapered-wing type. I t is emphasized that rectangular wings
do not have uniform loading, for reasons that will be brought out later.

186

BASIC WllQG A N D AIRFOIL TIfEORY

9.3. The Biot-Savart Law. Before \ye can actually calculate the
effect of the trailing vortices, it is necessary to develop vortex laws for
three-dimensional flow. Added value may be attached to this particular
development as i t embraces nearly every aerodynamic and vector principle studied so far.
First let us recall some of the rules and definitions of the velocity
potential. 4 has been defined by t,he relation

\\-herethe symbols are shown in Fig. 4.11.


The velocity potential of one point A relative to another point A' is

I t was demonstrated in Sect. 4.6 that 4 exists only for irrotational


motion and is then independent of
t,he path of integration-in
this
case AA' in Fig. 9.5. It is further
noted in Chap. 4 that the value of
the circulation bears a close resemblance to Eq. (9.3):

uAa

= $q.ds

(9.4)

The only difference is that for


circulation the path of integration
is closed. Under these conditions,
FIG.9.5.
if the line integral about the closed
curve is not zero, its value is the circulation, and the motion is not
completely irrotational unless a solid body is included.
Now consider an arbitrary surface which has a vortex for a boundary.
The line integral along any closed path 1M that does not cut the surface
is zero. But when the surface is cut, we get a sudden rise in potential
due t o the vortex. It is advantageous to assume a thickness to the surface to avoid an infinite rate of rise (see Fig. 9.6).
Now a given potential may be produced by sources and sinks or by
vortices. At this point it is more convenient to consider the flow field
produced by the closed vortex line of Fig. 9.5 as being produced by a
double sheet of sources and sinks as shown in Fig. 9.7. For instance,
if Q is the source quantity, then the velocity a t point P a distance r
A

187

THE F I N I T E W I N G

away will be
q = B
4nr2

(9.5)

[Compare Eq. (9.5) with the development of Eq. (3.13).]


sheet

Potential

The potential a t point a due to a source at the origin is

which in this case is

But q, is straight (the flow from a


source is radial), and the angle between q, and dr is zero (Fig. 9.8), so
that,
4,

==

qr dr

42w2

--

47ra

(9.8)

or for a sink

4a = 4na
If we call the source intensity Q1,
the potential at P due to elemental
area d A of the doublet is

FIG.9.8.

188

BASIC WING AND AIRFOIL THEORY

where a, and as are the distances from P to the proper source and sink
elements.
We see, however (Fig. 9.9), that
where hl is the distance between
source and sink sheets, and hence

Excluding the last term of Eq. (9.1 1)


as small, we have

We may drop the subscript on the a2 term as with hl small and a >> hl,
aig aa,and in the term where a2 appears the difference between ai and
a, is inconsequential.
Equation (9.9) is then

Now consider the flow between upper and lower sheets. Except for a
very small (but important) amount, the entire flow is from source to
sink between the sheets. For a square-foot area the quantity is Q1cu f t
per sec, and hence the velocity developed from upper to lower sheet is
numerically equal t o Q1f t per sec. 4 s this takes place along a distance
hl, we see the rise in potential = velocity X distance = Qlhl.
As we have defined the flow as producing a rise in potential equal to
bhe vortex strength (or I?), then Qlhl = I' and Eq. (9.13) becomes

Equation (9.14) relates the velocity potential with the circulation but
is not yet in a usable form. Indeed, from a practical standpoint, we
shall want to change it to another general form.
Now, return t,o the vortex-sheet concept, and put a small unit spherical
surface about the point P (Fig. 9.10a). From Fig. 9.10b the projection of
d A perpendicular to a is essentially d A cos a, and, assuming d A cos a
to be square, the rays from d A to P will define a small "square" (or

THE FINITE WING

189

solid angle w ) on the surface of the sphere. On the unit sphere the projection will be inversely proportional to the square of the distance, or

d A cos a

dw = a2

From Eq. (9.14)

or for the whole surface bounded by the vortex

Now the velocity a t point P is

We seek q due t o the vortex of strength I', and our development so far
tells us that q is a function of the solid angle w. I t is apparent that w
would change if the point P were moved from the surface or, equally,
if the vortex surface were moved from the point. The latter approach,
it develops, is easiest to analyze. We must move the sheet along a, hold-

190

BASIC JPI'lrC A S D A I R F O I L T H E O R )

ing it parallel to itself, as any angular motion vould correspond to rotating


the point P about the sheet, and v e n-ish to examine the effect of only
moving P.
From inspection, increasing the distance a decreases the potential a t P
as it decreases w. Thus it is that 17:hile we consider the effect of moving
the vortex sheet we are also considering the effect on the potential a t P.
From Fig. 9.11, the only movement of the vortex sheet that affects w
will be that which affect,sthe projected area d A 1 = d A cos a. Moving

the sheet parallel to itself, we see the net result is a decrease of o by a


perimetric strip, and we may evaluate this strip as follo~vs:The displacement along a of the vortex sheet parallel to itself creates a surface of
height dr and perimeter $ ds. A section of this surface is ds dr, and the
projection of ds dr on the perpendicular to a is ds dr cot 0 2 , where O2 is
the angle between the radius vector PL and a perpendicular to a, BN.
Now dr is perpendicular to ds, and the vector that represents the arca
ds dr may hence be written e ds dr or ds x dr. The vector that represents the area dAl will be perpendicular to dAl and hence lie along a
and have a unit length a l = a/a. In addition the angle O2 will be nearly
90, and we may write cot O2 = cos 02. Finally the change in projected
area as the sheet is moved becomes

T H E FINITE WING

and the change in the solid angle w is then

a3

(9.20)

which by Eq. (1.25) becomes

As we have moved the vortex sheet parallel to itself only, the direction
and length of dr is constant. Hence we may remove dr from the integral
[Eq. (9.21)] as follows:

fdT,

du = dr

(9.22)

To digress a moment,

From the definition of a derivative we then have


dr V w = dw

So from Eq. (9.23)

dr-Vw

d r - f y

and
Substituting in Eq. (9.18),

and substituting for d s x a its equal, ca ds sin 0, where 0 is the angle


between a and the vortex segment ds, we have (see Fig. 9.126)
191 =

&/,

,
ds sin 9

Now, having the expression for the effect of a segment of vortex, we


may proceed to the application problems. First, however, it is convenient to express Eq. (9.25) in trigonometric terms.
Taking out a section a dB (Fig. 9.12a), we see that d x = a dB;
sin B = d x / d s
so that

a dB
d~ = sin 0

192

BASIC WING AND AIRFOIL THEORY

Also from Fig. 9.12b sin 0 = h/a, so that

Changing limits, when s = A, 0


stituting, Eq. (9.25) becomes

01 and when s = B, 0 =

02,

and, sub-

I'

lql = - (cos 02 - cos 01)


4rh

Considering the two interior angles as easier to use, we have

Equation (9.27) is a common form of the Biot-Savart law. I t demonstrates first of all that, as the vortex segment is approached, the induced

velocity it produces rapidly increases. Now, according to the Helmholtz


laws the vortex cannot abruptly end in the fluid, but when it turns a
sharp corner the usual approach is to consider each straight segment
separately, just as though it did terminate at the corner. Thus for a
vortex that "starts" at P and extends to infinity out beyond B, el = r j 2 ,
and 83 = 0, and

9 "

(9.28)

q,h.

Or if the vortex extends to infinity in both directions,

el

e3 = 0 and

Of interest is the fact that the induced velocity exists beyond the
"end" of a vortex, although its value is small (Fig. 9.13).
The replacing of a wing by a simple lifting-line vortex and the proper
use of the Biot-Savart law are illustrated in subsequent chapters.

193

THE FINITE WING

9.4. Downwash at the Wing for Any Spanwise Distribution of Circulation. Undoubtedly the most important effect of the trailing vortices

FIG.9.13. Values of induced velocity near the "end" of a vortex segment, I'
It per sec.

=400~
sq

is the increase of drag they produce through altering the flow angle a t
the wing. Secondarily, the manner in which they alter the flow at the
wing is important, too, as usually the induced velocity is not constant
across the span, and the local angle
of attack varies over a large range.
Let us call the downwash velocity w and the wing span b. Orient
the wing so that the x axis coincides with the forward airplane
longitudinal axis, the y axis coin- 1:
cides with the bound, or lifting,
vortex, and the z axis is downward.
(Downwash is positive in the downward direction.)
Yz
Since the circulation decreases toFro. 9.14.
ward the tips (it must be zero at the
tip), between the points y and y
dy on the span the circulation changes
+--

- ar
-dy.
a~

-------

Hence the induced velocity a t the point yl for a differential

vortex a t y is, from Eq. (9.28) and Fig. 9.14,

and the total downwash a t y l will be

* dw(yt) should be read "dw at yl."

191

BASIC WING A N D AIRFOIL THEORY

Equation (9.31) hence defines the amount of downwash that will occur
at point yl for any arbitrary distribution of circulation described by
dr
-.
8~

I t is known as the Prandtl downwash relation.

Using the local down~vashw, ~i-emay then find the local angle of
attack and local lift and finally the approximate span load distribution.
Repeating the process, we could in theory find the actual span loading
by eliminating the effect of each new downwash distribution on the
original assumed span loading. However, the problem is extremely
complex, and we shall usually confine ourselves to considering a uniform
or an elliptic span loading in this and the following chapter. (In Chap. 12
two methods for getting the span loading for arbitrary planforms are
discussed.)
9.5. Downwash at the Wing for Elliptic Distribution of Circulation.
Before proceeding further it is in order to discuss the terms span loading,
spanwise lift distribution, and spanwise lift coeficient distribution. Since
[Eq. (4.4)]the lift is a function only
of the density, velocity, and circulation, for a given set of conditions
an elliptic distribution of circulation would also be an elliptic distribution of lift. However, the relation between the section lift and
the section lift coefficient involves
no.9.15.
the local chord, and hence elliptic
lift or distribution of circulation may or may not mean elliptic distribution of cl. I n fact, the only case where all three distributions are elliptic
is that of the properly twisted rectangular wing.
Proceeding now to the case of elliptic loading ("loading" as we have
seen may be either lift or circulation), we assume a wing that has a circulation roat mid-span which decreases elliptically until it is zero a t the
wing tips (see Fig. 9.15). We write

or

Then

THE FINITE WING

195

Substituting into Eq. (9.31), we get that the downwash a t point yl on the
span is

The integrations of Eq. (9.35) is complicated by the fact that the value
of the integrand is infinite when y = yl, b/2 or -b/2.
The complete
integration is given in Appendix 3, and we may hence state

But yl is any point on the span, and hence this relation applies to every
point on the span.

The interesting conclusion'is that, if the distribution of circulation is


elliptic across the span, the downwash a t the wing due to the trailing
vortices is constant across the span. While we might rightfully conclude that the local angle of attack is also constant for an untwisted
elliptic wing, the conclusion cannot be extended to include a simultaneous
stall across the span, as practical considerations of viscosity dictate an
earlier stall for the shorter chord, lower Reynolds number wing tips.
9.6. Downwash at the Wing for Uniform Distribution of Circulation.
The case of uniform distribution of circulation is developed in Chap. 11,
but we may state the results here to compare with Sect. 9.4. It is shown
in Chap. 11 that the downwash distribution a t a wing with uniform
loading is

(The symbols are defined in Chap. 11.)


Substituting for example r = 6.28 and s = 1.0, we get the values of w
shown in Fig. 9.16. Here it is seen that the angle of attack is locally
reduced by uniform loading all along the span, but unlike the case of
elliptic loading the reduction is not uniform, being greatest at the tips.
This tends toward a wing root stall appearing earliest. We may also
get nearly uniform loading by twisting a tapered wing, but in practice
the twist is kept small to avoid excessive deviations from the elliptictype loading, for reasons to be advanced later.
As a matter of interest the effect of the trailing vortices on the uniformly loaded wing quite obviously tends toward destroying its uniformity of lift, and hence even the rectangular wing tends toward an

BASIC WING AND AIRFOIL THEORI'

-M

197

THE FINITE WING

elliptic distribution. Experience tells us it gets about halfway there,


but the alteration from the basic uniformity is not enough to alter the
expected root stall.
An interesting check of this theory and of Fig. 9.16 was made by
Piercy and is reported in Ref. 9.3. In these experiments a wing of 12 in.
span and 3 in. chord was set at 8" in a wind tunnel with a 4-ft-square test
section. Using a small yawhead, the angle of flow was measured 0.9b
behind the wing in a plane near that of the wing. The results, shown in
Fig. 9.17, demonstrate clearly the existence of a vortex similar to that
supposed by the theory. At a distance of 0 . 5 ~behind the trailing edge
a trailing vortex was clearly discerned. Its core had a diameter of 0 . 2 ~ ~
and its center was 0.95s from the plane of symmetry. Inside the core
the linear velocity appeared very nearly proportional to the radius (in
accordance with Fig. 4.3), and outside the core the velocity closely
checked the values of Eq. (9.37). The vortex span was somewhat greater
than predicted in Sect. 11.6 but remained constant with a.
The diameter of the vortex core (for which no satisfactory theory has
been devised) increased slightly with the circulation. The center of the
core for this particular case revolved a t over 18,000 rpm.
The high speed of the vortex core produces interesting and beautiful
effects in air under special conditions. The high speed implies a low
pressure, and for air we then get a low temperature. Since the amount
of moisture the air can hold decreases with falling temperature, vortices
in humid air becomes visible, appearing somewhat like steam in air, as
the moisture condenses out. This phenomenon is a familiar sight a t the
wing tips and landing-flap terminations and, more frequently, streaming
helically behind propeller blade tips a t take-off.
9.7. Induced Drag Coefficient. We may now examine the effect of
the downwash on drag for the simplest case-that of elliptic loading.
For the finite wing, Eq. (4.25) becomes

and for elliptic loading in accordance with Eq. (9.33) we get

Making the substitution that y = (b/2) sin 0, dy


finding new limits we have

(b/2) cos e de and

198

BASIC WING AND AIRFOIL THEORY

Integrating and solving for Fo, we have

Substituting into Eq. (9.36),

and it is seen that the effect of downwash is to deflect the airstream


through an angle
W

(9.43)

a, = -

as is shown in Fig. 9.18. The changes in lift and velocity are small and
may be neglected, but the effect on
drag may be large.
Considering the dragwise component
of the lift Di,we have
L

D,

(3

L tan a; = L -

(9.44)

and from Eq. (9.43) this becomes


L2

D, = (p/2) V 2rb2

(9.45)

FIG.9.18.

Multiplying both sides of the equation by V and dividing by 550 to get horsepower we have, since the lift
must equal the weight W in steady unaccelerated flight,

where q = dynamic pressure (p/2)V2. Putting V in miles per hour and


letting a = p/po, where po = density at standard conditions, the final
form is

Equation (9.47) demonstrates that the horsepower required to overcome induced drag is a function of the s p a n loading, the velocity, and the
density and gives a clearer insight into the problem than the usual

CL*
CDi = T A.R.

T H E FINITE W I N G

199

which may be found from Eq. (9.45). Equation (9.48) seems t o indicate that induced drag is a function of aspect ratio, where, more accurately, only the drag coefiient varies with aspect ratio, and the area
changes t h a t occur with changing aspect ratio and constant span are
such that. the induced drag remains constant with constant weight and
speed.
It should be noted that Eqs. (9.45), (9.47), and (9.48) were derived
for elliptic loading only.
PROBLEMS

9.1. If a wing of 6 f t chord has a local lift coefficient of 0.6 and is operating in a
100-mph airstream, what is the local circulation?
9.2. Prove that the untwisted elliptic wing gives elliptic distribution of circulation.
9.3. What is the downwash in feet per second at the wing centerline of a wing
with uniform loading if its lift is 10,000 Ib, span 30 ft, V = 200 mph, and the air
standard sea-level density?
9.4. Find the induced velocity 4 ft from a doubly infinite vortex of strength
r = 30 sq f t per sec.
REFERENCES

9.1. F. W. Lanchester, "Aerodynamics," Constable & Co., Ltd., London, 1918.


9.2. L. Prandtl, Applications of Modern Hydrodynamics to Aeronautics, TR 116,
1921.
9.3. N. A. V. Piercy, On the Vortex Pair Quickly Formed by Some Airfoils,
JRAS, October, 1923, p. 489.
9.4. S. Goldstein, Modern Developments in Fluid Dynamics, Vol. I, p. 69,
Oxford University Press, New York, 1938.
9.5. N. A. V. Piercy, "Aerodynamics," p. 235, English University Press, London.

Potrebbero piacerti anche