You are on page 1of 17

JOURNAL OF OPTIMIZATION THEORY AND APPLICATIONS: Vol. 5, No.

6, 1970
Hypersonic Airfoils of Maximum lift-to-Drag Ratio!
R. A. THOMPSON
2
AND D. G. HULL
3
Communicated by A. Miele
Abstract. The problem of determining the slender, hypersonic airfoil shape
which produces the maximum lift-to-drag ratio for a given profile area, chord,
and free-stream conditions is considered. For the estimation of the lift and
the drag, the pressure distribution on a surface which sees the flow is
approximated by the tangent-wedge relation. On the other hand, for surfaces
which do not see the flow, the Prandtl-Meyer relation is used. Finally, base
drag is neglected, while the skin-friction coefficient is assumed to be a
constant, average value. The method used to determine the optimum
upper and lower surfaces is the calculus of variations. Depending on the
value of the governing parameter, the optimum airfoil shapes are found to be
of three types. For low values of the governing parameter, the optimum
shape is a flat plate at an angle of attack followed by slightly concave upper
and lower surfaces. The next type of solution has a finite thickness over the
entire chord with the upper surface inclined so that the flow is an expansion.
Finally, for the last type of solution, the upper surface begins with a portion
which sees the flow and is followed by an inclined portion similar to that
above. For all of these solutions, the lower surface sees the flow. Results are
presented for the optimum dimensionless airfoil shape, its dimensions,
and the maximum lift-to-drag ratio. To calculate an actual airfoil shape
requires an iteration procedure due to the assumption on the skin-friction
coefficient. However, simple results can be obtained by assuming an ap-
proximate value for the skin-friction coefficient.
1 Paper received November 26, 1969. This research was supported in part by the Air Force
Office of Scientific Research, Office of Aerospace Research, U.S. Air Force, under AFOSR
Grant No. 69-1744.
2 Graduate Student of Aerospace Engineering, University of Texas, Austin, Texas.
3 Assistant Professor of Aerospace Engineering, University of Texas, Austin, Texas.
432
]OTA: VOL. 5, NO.6, 1970
433
1. Introduction
Previous analyses of optimum hypersonic airfoil shapes have been con-
cerned with nonlifting, symmetric airfoils of minimum drag or lifting,
flat-top airfoils of minimum drag for a given lift or of maximum lift-to-drag
ratio. The former are summarized in Ref. 1, while the latter are contained in
Refs. 2-7. In this paper, lifting airfoils are optimized, but the only restriction
imposed on the upper and lower airfoil surfaces is that their slopes relative to
the free-stream direction are small, implying that the airfoils are slender and
are at small angles of attack.
If such conditions as extreme heat-transfer rates, use of the wing as a
storage volume, or the necessity to maintain structural strength are not con-
sidered, the lift-to-drag ratio is the primary indicator of wing performance.
However, it is easily shown that a wing designed solely to maximize the lift-to-
drag ratio is too thin for practical purposes, at least from the standpoint of
internal volume or structural strength (for example, see Ref. 6, where thickness
ratios on the order of 2 % are predicted). Therefore, it is desired to determine
the greatest performance that can be achieved when the prescribed characteris-
tics of the wing are considerably different from those of the unconstrained
optimum wing (Ref. 6) when both are operating at a specified flight condition,
that is, altitude and velocity.
In view of the above statement, the problem of determining the airfoil
shape which maximizes the lift-to-drag ratio for a given profile area and chord
is considered here. It is assumed that the pressure distribution on surfaces
which see the flow can be approximated by the tangent-wedge relation. On
surfaces which do not see the flow, the Prandtl-Meyer pressure relation for
small disturbances in hypersonic flow is employed. With regard to skin
friction, it is assumed that the surface-averaged skin-friction coefficient can
be held constant during the optimization procedure. It has been shown in
Ref. 7 that this assumption is valid and that it simplifies the extremization
problem considerably. Finally, the base drag is neglected with respect to the
forewing drag. The effect of including the base drag in optimum hypersonic
airfoil studies is negligible for moderate to high hypersonic Mach numbers
(Ref. 7).
2. Lift-to-Drag Ratio and Profile Area
In order to determine an optimum airfoil shape, it is necessary to predict
the appearance of the optimum airfoil shape. Now, the family of optimum
434
lOTA: VOL. 5, NO.6, 1970
airfoil shapes obtained by specifying the profile area and the chord must
contain the family of optimum airfoil shapes obtained by specifying only the
chord. Next, the optimum airfoil shape for a given chord is a wedge at such an
angle of attack that the upper surface is located in an expansion region of the
flow field (Ref. 6). Therefore, to begin the solution of the present problem, it
is assumed that the optimum airfoil shape has the form shown on the left
in Fig. 1.
If it is assumed that the airfoil is slender and is operating at a small angle
of attack, that the effect of skin-friction forces on the lift is negligible with
respect to pressure forces, that the base drag is negligible with respect to the
forebody drag, and that the surface-averaged skin-friction coefficient C fa is
constant during the optimization procedure, the expressions for the lift L
and the drag D can be written as
Ljq = f: (C
p
! - Cpu) dg
(1)
Djq = 5: (Cp!TJ!' - CpuTJu' + 2C
ta
) dg
where q is the free-stream dynamic pressure, C
p
is the pressure coefficient,
u and I refer to the upper and lower surfaces, respectively, and r/ denotes the
derivative dry/dg. It should be noted that c, the length being held constant, is
the chord under the assumptions of slender airfoils at small angles of attack.
In these expressions, the pressure coefficient on the upper surface is given by
the Prandtl-Meyer expansion formula, while the pressure coefficient on the
lower surface is obtained from the tangent-wedge approximation (Ref. 8),
that is,
c
~
!
x
T
T
t
MT
1
1
7J
y,z
Fig. 1. Coordinate systems.
}OTA: VOL. 5, NO.6, 1970 435
where y is the ratio of specific heats, M is the free-stream Mach number, and
ki = (y - 1)/2, k2 = 2y/(y - 1), k3 = (y + 1)/4. Finally, the profile area A
is given by
(3)
At this point, the following dimensionless variables and parameters are
introduced (Fig. I-right):
x = ~ / c ,
E* =L/MD,
as are the functions
y = M'1]u/
c
,
A* = MA/c
2
,
Then, the lift-to-drag ratio can be written as
where
while the profile area becomes
1
A*=fo(z-y)dX
3. Maximum Lift-to-Drag Ratio Problem
T = tic
M* = M {leta
(4)
(5)
(6)
(7)
(8)
The problem of maximizing the lift-to-drag ratio for a given profile area
and chord is equivalent to that of maximizing the ratio of integrals (6) subject
to the isoperimetric constraint (8) and certain prescribed boundary conditions.
This is not a standard problem of the calculus of variations, but it can be
transformed into one (Ref. 9). In other words, it can easily be shown that the
problem of maximizing the ratio of functionals E * = I L/I D is equivalent to
that of maximizing the modified functional E' = IL - E*I
D
. However, it is
known a priori that the maximum value of E' is zero, that is, E* = IL/ID ,
and that E* is actually the maximum value being sought and, hence, is held
constant during the extremization procedure. To constrain the profile area,
436 JOTA: VOL. 5, NO.6, 1970
Eq. (8) is adjoined to the modified functional E' by an undetermined, constant
Lagrange multiplier A. The proof of these statements can be conducted by
reformulating the above problem as a Mayer problem or by variational differen-
tiation.
The result of the above discussion is that the problem of finding the
airfoil shape of given profile area and chord which maximizes the lift-to-drag
ratio is equivalent to that of maximizing the standard-form functional
XI
1= J F(y, z,y, i, E*, A) dx
Xi
(9)
with
F = (yQ - P) - E*(yiQ - yP + yM*3) + A(Z - Y) (10)
subject to the prescribed boundary conditions
Xi = 0, Yi = 0,
YI free,
Zi = 0
ZI = free
(11)
XI = 1,
and the integral relations (6) and (8) which are used to determine E* and A as
functions of the profile-area parameter A * .
4. Necessary Conditions and Solution Process
It is known that the functions y(x) and z(x) which extremize the functional
(9) must be solutions of the Euler equations (see Ref. 1, for example)
Fy - dFy/dx = 0, F
z
- dFi/dx = 0 (12)
which, since Fy = -A and F
z
= A from Eq. (10), admit the first integrals
F 11 = -Ax + C
1
, (13)
The constants of integration C
1
and C
2
can be determined by applying the
transversality. condition
[(F - yFy - iF
i
) ox +Fy oY +F
t
oz]r = 0 (14)
which must be consistent with the prescribed boundary conditions. Since
8y j *- 0 and 8z j *- 0 while the remaining variations at the end points vanish,
the transversality condition requires that (Fy)j = (Fi)j = 0, so that C
1
= A
}OTA: VOL. 5, NO.6, 1970
437
and C
2
= -,\ from Eq. (13). Finally, taking the required derivatives of the
fundamental function (10) leads to the following equations for determining
the optimum upper and lower airfoil surfaces:
P' - E*(P + yP') = -A(1 - x), yQ' - E*(yQ + yzQ') = -A(1 - x)
(15)
where P' = dPjdy and Q' dQjdz.
The procedure employed to obtain the functions y(x) and z(x) which
maximize Eq. (9) is an iterative one. Values are chosen for y and M*, and
values are guessed for the constants E* and ,\. Then, Eqs. (15) are solved for
y(x) and z(x) by incrementing x in the interval 0 :( x :( 1. These values of
the derivatives are then used in Eq. (6) to calculate a new value of E* . This
value of E* differs from the guessed value, and the difference can be denoted
by ,1E*. What remains is to iterate on the guessed E* until ,1E* = 0
(,1E* :( 10-
7
, in this case), and this is done by the Newton-Raphson method.
Once the correct value (that is, the maximum value of E*) is determined, the
final set of slopes y(x), z(x) is integrated by Simpson's rule to obtain
the optimum airfoil shape y(x), z(x). Finally, the thickness ratio parameter
T* = Zj - Yj can be determined, and the integration indicated in Eq. (8)
2.0 r---.::::-r---..:---.,----...,----..,..---.,----...,----..,..,,-----,
o
-1.0
-2D 1---+-----1----- -- -
___ __ __ ___ __ __ __
-0.1 o 0.1 .. 0_2
Y,z
0.3
Fig. 2. Euler equations.
0.4 0.5 0.6
438
}OTA: VOL. 5, NO.6, 1970
can be carried out to find the value of A* corresponding to the prescribed
value of A.
A word as to how E* and A are guessed is in order. Hypersonic airfoils can
be shown to develop lift-to-drag ratios on the order of five. Hence, at a Mach
number of five, E* would have a value of unity, and one might make the first
guess E* = 1. With regard to possible choices of A, Eqs. (15) are plotted in
Fig. 2 for y = 1.4 and various values of E* , and only that part of the graph
where y, z are positive pertains to the discussion. At the origin, where x = 0,
the ordinate of Fig. 2 becomes just A. Hence, for E* = 1, one could choose
any value of A in the range -0.02 ::::;;; A ::::;;; 1.4 and obtain a solution for which
Zi ~ Yi and Yi ~ O. By working around these values, the complete set of
solutions for which Y, Z are positive can be found.
In carrying out these calculations, the ratio of specific heats is chosen to
be 'Y = 1.4, and the parameter M * is varied in the range 0.5 ::::;;; M * ::::;;; 1. 5.
The lower limit defines the beginning of hypersonic flow, while the upper
limit defines that part of hypersonic flow where viscous interactions become
important. The optimum shapes are shown in Figs. 3-5 with A as a parameter.
The relation between A and A* (the parameter containing all the prescribed
quantities) is shown in Fig. 6. Finally, the approximate thickness-ratio para-
-1.0
A =6
y
o
---
3
1.4
~
-----=
-CD
------
1.4
~
--
3
1.0
M.=0.5
' ~
~
~
z
2.0
3.0
o 0.2 0.4 0.6 0.8 1.0
x
Fig. 3. Optimum airfoil shapes, M* = 0.5.
JOTA: VOL. 5, NO.6, 1970 439
-1.0
Y
A=
3
0
1.4
0
-CD
1.0
z
2.0
3.0
0 0.2 0.4 0.6 0.8 1.0
x
Fig. 4. Optimum airfoil shapes, M* = 1.0.
-1.0
Y A=6
3
0
z

3.0
0

0.6 0.8 1.0
x
Fig. 5. Optimum airfoil shapes, M* = 1.5.
440
IOTA: VOL. 5, NO.6, 1970
8 r---------,---------,---------,---------,--------,
6 r - - - - - - - - - T - - - - - - - - - r - - - - - - - - - r - - - - - - - ~ ~ ~ ~ - - ~
2
o
-2
o 0.4 0.8 1.2 1.6 2.0
Fig. 6. Lagrange multiplier.
~ O r - - - - - - - - - - r - - - - - - - - - - . - - - - - - - - - , - - - - - - - - - - . - - - - - - - - - - ,
4.01--------+--------1------+------+--------1
Aft
Fig. 7. Thickness-ratio parameter.
IOTA: VOL. 5, NO.6, 1970 441
meter 'T * is shown in Fig. 7, while the maximum lift-to-drag ratio parameter
E* is plotted in Fig. 8.
Theoretically, all the arcs obtained so far are extremal arcs. To ascertain
that these extremal arcs are indeed maximal arcs, one must show that the
Legendre condition
(16)
is satisfied at each point of each extremal arc. Because of the nature of the
problem, FYi = 0, and the Legendre condition is satisfied if
0,
(17)
at each point. The explicit forms of these inequalities are obtained by using
Eq. (10) and are given by
_pH + E*(2P' + yP") 0, Q" - E*(2Q' + zQ") (18)
and are plotted in Fig. 9 for several values of E * . Since the Legendre condition
is satisfied, the extremal solutions are maxima.
The solutions obtained in this section are only valid for a finite range of
22
14

!

/Yj=O
I
!
1

i
i
\.
VYj=i
j
i \



Y -
". 1.0 t'-.....
\
\
'-.
""-
\
' ......
--
1.8
10
0.6
1.5
r--._
0.2
o 0.4 0.8 1.2
A.
Fig. 8. Lift-to-drag ratio parameter.
1.6 2.0
442
JOTA: VOL. 5, NO.6, 1970

-5 __
F"
yy
-10 ---.-
____ ______ L-______ ______ ____ ____
-2.0 -1.0 o
Y,Z
1.0 2.0 3.0 4.0
Fig. 9. Legendre condition.
values of ,\. One limiting value, ,\ = y, occurs when Yi = 0; for ,\ > y, the
initial slope of the upper surface becomes negative, creating a compression
rather than an expansion for which the pressure law has been chosen. However,
the expansion formula holds for the compression provided Y is very small
(Ref. to). At any rate, for ,\ > y, the problem is reformulated to include the
compression on the upper surface, and this is discussed in the following section.
The other limiting value occurs when Yi = Zi ; however, the value of ,\ at
which this occurs varies with M * . For values of ,\ lower than this limiting
value, Eqs. (15) predict that Yi > Zi' which is not physically possible. To
find these solutions, it is necessary to add whatever constraints are necessary to
ensure that the two surfaces do not intersect. It is observed that this class of
solutions must contain that for A* = 0 (Fig. 5), which is a flat plate (zero
profile area) at the optimum angle of attack. Finally, it is observed that the
optimum shapes for the values of ,\ being considered are composed of a flat
plate followed by concave upper and lower surfaces which give the airfoil
thickness near the trailing edge. These shapes have not been computed in
their entirety because they are not practical. However, the correct trends are
shown in the figures by dashed lines. The points for A * = 0 (,\ = - (0) have
been calculated.
]OTA: VOL. 5, NO.6, 1970
443
5. Necessary Conditions and Solution Process, A> Y
From the trend of the solutions for'\ ~ y, it is apparent that the solutions
for ,\ > y have a similar lower surface, but the upper surface begins with a
negative-slope portion (tangent-wedge pressure distribution) and switches to
a positive-slope portion (Prandtl-Meyer pressure distribution) at some abscissa
Xc to be determined during the optimization procedure. For these shapes,
the lift-to-drag ratio is given by
(19)
where
IL = {C- (yQ - yR) dx + f (yQ - P)dx
o Xc+
(20)
ID = {C- (yzQ - yjR + yM*3) dx + f (yzQ - yP + yM*3) dx
o Xc+
and where
(21)
In view of the discussion at the beginning of the previous section, the
problem of maximizing the lift-to-drag ratio (19) for a given profile area (8) and
chord is equivalent to the problem of maximizing the functional
with
f
"'C- fl
I = G(y, z, y, Z, E* , ..\) dx + F(y, z, y, Z, E* , ..\) dx
o Xc+
G = (yQ - yR) - E*(yzQ - yjR + yM*3) + ..\(z - y)
F = (yQ - P) - EAyzQ - yP + yM*3) + ..\(z - y)
subject to the prescribed boundary conditions
XI = 1,
Yi = 0,
YI = free,
Zi = 0
Zl == free
(22)
(23)
(24)
and the integral relations (8) and (19) which are used to determine E* and A as
functions of the area parameter A * .
444
JOTA: VOL. 5, NO.6, 1970
The functional (22) is not one of the standard functionals for which
necessary conditions have been derived. It is therefore necessary to develop
the necessary conditions through variational differentiation. Hence, after
taking the first variation of Eq. (22), performing the usual integration by parts,
and relating variation 8 taken at constant x to arbitrary variations S (see Ref.1,
for example), one obtains the following relation:
OJ = (C- [(Gy - dG,iJ/dx) 8y + (G
z
- dGz/dx) 8z] dx
"'.
+ [(G - yGy - zG
z
) 8x + Gy 8y + G
i
8z]f-
+ (' [(Fy - dFiI/dx) 8y + (Fz - dFt/dx) 8z] dx
"'c+
(25)
This relation does not allow for a corner except at the point Xc ; however, if a
corner exists on either the upper or lower surface between Xi and Xc or Xc
and xf ' then the standard corner conditions apply. The vanishing of the first
variation (25) leads to the Euler equations
Gy - dGy/dx = 0,
Fy - d } ~ / d x = 0,
G
z
- dGi/dx = 0,
F
z
- dFi/dx = 0,
(26)
the natural boundary conditions (SXi = SYi = SZi = SX
f
= 0, SYt * 0,
SZf * 0)
(27)
(G - yGy - zGi)c_ = (F - yFy - zFt)c+
(Gy)c_ = (Fy)c+ , (Gz)c- = {Fz)c+
(28)
In applying these conditions, there is no reason to suspect a corner on
the lower surface. On the other hand, the corner conditions must be applied
to the upper surface to determine the conditions under which the compression
and expansion segments are to be joined. In this connection, if zc_ = Z'c+ ,
then the last corner condition is satisfied since G
z
= Fi from Eqs. (23). With
this information, the remaining corner conditions reduce to
(29)
F
Y
6
4
2
o
-2
20
16
12
8
4
JOTA: VOL. 5, NO.6, 1970
1\ \ \
\ \\
( Fyle_
' 0 ~
~
I
~
~ . = O
r-----I-----
\:
r- 1.0
2.0
-3 -2 -I
'Ie -
o 2
Ye+
Fig. 10. Corner condition.
\E.=2.0
\
V
G-yG
y
\.
'"
'--Z
F-yF
y
o
-1.2
~ ~
-0.8 -0.4
o 0.4 0.8
Yc-
Yc+
Fig. 11. Corner condition.
445
(Fyle+
3 4
r2.0
Jf'O
E.=O
1.2 1.6
446
IOTA: VOL. 5, NO.6, 1970
and, after the use of Eqs. (23), are plotted in Figs. 10-11. These figures show
that the only way the compression and expansion subarcs can be joined so
that E * is the same for each subarc is if
Yc- = Yc+ = 0
(30)
Next, if it is observed that G
y
= -A, G
z
= A, Fy = -A, and F
z
= A from
Eqs. (23), the Euler equations (26) can be integrated to obtain
G
y
= -Ax + C
I
,
Fy = -Ax + C
3
,
(31)
Hence, the last two corner conditions in Eq. (28) require that C
1
= C
a
and
C
2
= C
4
, while the natural boundary conditions (27) show that C
a
= A and
C
4
= -A. Combining these results with the first integrals (31) and the
fundamental functions (23) leads to the following equations for determining
the optimum airfoil shape:
yR' - E*(yR + yjR') = -A(1 - x)l
yQ' - E*(yQ + yzQ') = -A(l - x)
P' - E*(P + yP') = -A(l - x)l
yQ' - E*(yQ + yzQ') = -A(l - x)!
(32)
Finally, the abscissa Xc where the upper surface subarcs are joined can be
determined from (Gy)c_ = -A(l - xc), Yc- = 0, so that (Gy)c_ = Y;
therefore,
Xc = 1 - yfA (33)
To show that the extremal arcs obtained above are maximal arcs, It IS
necessary to show that the second variation is nonpositive. For fixed end points,
the second variation becomes
'iN = {C- [G
yy
(&y)2 + G
U
(&Z)2] dx + {f [Fyy(&y)2 + Fu(8z)2] dx ~ 0 (34)
:Ilj :Ilc+
This condition is satisfied if
Gyy ~ 0,
Fyy :< 0,
G u ~ O ,
F u ~ O ,
Xi ~ X < Xc
where, from Eqs. (23), it is seen that GJi = FJi .
(35)
JOTA: VOL. 5, NO.6, 1970
447
The procedure to be followed in the calculation of the airfoil shapes is
nearly identical with that used previously. The difference is that the calculation
is started using the first two of Eqs. (32). When the upper-surface slope y
becomes zero, the equations must be changed to the last two of Eqs. (32).
The selection of the values for E * is essentially determined by where the
previous solutions left off. However, the choice of values for A is limited to the
values A ~ y for which Yi ~ 0 (Fig. 2). The optimum shapes are shown in
Figs. 3-5; the quantities A, T*, E* are plotted versus A* in Figs. 5-8; and the
satisfaction of the Legendre condition is displayed in Fig. 9.
6. Discussion and Conclusions
The problem of determining the airfoil shape which produces the
maximum lift-to-drag ratio for a given profile area and chord has been consid-
ered. It is found that the airfoil shapes fall into three categories, depending
on the value of the area parameter A* . For low values of A* , the optimum
shape is composed of a flat plate at an angle of attack followed by concave
upper and lower surfaces which give the airfoil thickness near the trailing edge.
For A* = 0, the optimum airfoil shape is a flat plate at the optimum angle of
attack. These airfoil shapes are considered impractical because of the low
thickness ratios. The airfoil shapes for moderate values of A* have a finite
thickness all along the chord and include the optimum shape for the case where
the profile area is free (A = 0), that is, the wedge. These airfoil shapes have
the highest maximum lift-to-drag ratios, but the thickness ratios may be too
low (around 2 %, see Ref. 6) for practical purposes. For the larger values of A* ,
the optimum airfoil shapes have a lower surface which always sees the flow,
while the forward portion of the upper surface sees the flow and the rearward
portion does not see the flow. For these shapes, the thickness-ratio parameter
increases with A* and the lift-to-drag ratio parameter decreases.
In order to determine actual airfoil shapes, it is necessary to employ an
iteration procedure. This procedure begins by specifying the free-stream
conditions and guessing a value for C ta . An optimum shape is then determined,
and the value of C
ta
for this shape is calculated and is used as the guess for the
next iteration. The procedure is repeated until the guessed and calculated
values agree. It is possible to obtain some approximate information by assuming
that the boundary layer is laminar and that C ta is for a flat plate at zero angle
of attack. If the flight conditions are characterized by a free-stream Mach
number of M = 10 and a free-stream Reynolds number of Re = 10
6
, the
value of C
ta
given in Ref. 11 is C
ta
= 10-
3
Using these numbers, one finds
448 ]OTA: VOL. 5, NO.6, 1970
that M * = 1, so that, if Ajc
2
= 0.05, the following results are obtained:
T = 0.084 andLjD = 6.7.
While this analysis has omitted the base drag as far as the optimization
problem is concerned, the calculation of the final lift-to-drag ratio must
include the base drag. Hence, the above value ofLjD decreases somewhat.
Finally, it is noted that the optimum airfoil shapes obtained here have
sharp leading edges and, hence, are subject to heating rates excessive for
present-day materials. The unavoidable blunting of the leading edge causes
a further decrease in the lift-to-drag ratio.
References
1. MIELE, A., Editor, Theory of Optimum Aerodynamic Shapes, Academic Press,
New York, 1965.
2. DONALDSON, C. DUP., and GRAY, D. E., Optimization of Airfoils for Hypersonic
Flight, Aerospace Engineering, Vol. 20, No.1, 1961.
3. LEWELLEN, W. S., and MIRELS, H., Optimum Lifting Bodies in Hypersonic Viscous
Flow, AlAA Journal, Vol. 4, No. 10, 1966.
4. MIELE, A., and LUSTY, A. H., Jr., On Optimum Wedges and Semicones in Hyper-
sonic Viscous Flow, AIAA Journal, Vol. 5, No.1, 1967.
5. HULL, D. G., Two-Dimensional, Lifting Wings of Minimum Drag in Hypersonic
Flow, Rice University, Aero-Astronautics Report No. 24, 1966.
6. HULL, D. G., Two-Dimensional, Hypersonic Wings of Maximum Lift-to-Drag
Ratio, Journal of the Astronautical Sciences, Vol. 14, No.2, 1967.
7. WEEKS, D. E., Jr., and HULL, D. G., Optimum Flat-Top Airfoils in Viscous
Hypersonic Flow, Journal of Optimization Theory and Applications, Vol. 5,
No.1, 1970.
8. Cox, R. N., and CRABTREE, L. F., Elements of Hypersonic Aerodynamics, Academic
Press, New York, 1965.
9. MIELE, A., On the Minimization of the Product of Powers of Several Integrals,
Journal of Optimization Theory and Applications, Vol. 1, No.1, 1967.
10. DORRANCE, W. H., Two-Dimensional Airfoils at Moderate Hypersonic Velocities,
Journal of the Aeronautical Sciences, Vol. 19, No.9, 1952.
11. SCHLICHTING, H., Boundary Layer Theory, Pergamon Press, New York, 1955.

You might also like