You are on page 1of 104

NASA

TechnicaI
Paper
2771
1988
3upersonic
Aerodynamics
of Delta Wings
Richard M. Wood
Langley Research Center
Hampton, Virginia

National Aeronautics
and Space Administration

Scientific and Technical


Information Division
Contents
Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

Symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
Aerodynamic and Geometric Parameters . . . . . . . . . . . . . . . . . . . . . . 3
Zero-Lift Wave Drag . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
Diamond airfoil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
Circular-arc airfoil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
airfoil . . . . . . . . . . . . . .
Modified four-digit-series . . . . . . . . . . . 5
Comments on zero-lift wave drag . . . . . . . . . . . . . . . . . . . . . . . . . 7
Lifting Characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
Lee-side flow characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
Flat delta wings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
Thickness effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
Camber effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
Comments on lifting characteristics . . . . . . . . . . . . . . . . . . . . . . . . 16

Real-Flow Wing Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

Appendix-Nonlinear Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . 20

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

Table . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
I

...
111
Summars design of aerodynamic vehicles. This change in re-
search emphasis was no doubt highly influenced by
Through the empirical correlation of experimen- the challenge of a new frontier-efficient supersonic
tal data and theoretical analysis, a set of graphs have flight.
been developed which summarize the inviscid aero- In the design of an optimum wing at super-
dynamics of delta wings at supersonic speeds. The sonic speeds, the aerodynamicist may select from
zero-lift wave-drag characteristics of delta wings with an infinite array of wing planforms, airfoil profiles,
diamond, circular-arc, and NACA modified four- and camber and twist distributions to meet the de-
digit-series airfoils were determined through the ap- sired aerodynamic performance requirements. His-
plication of a nonlinear computational technique. torically, the aerodynamicist relied heavily on linear
The nonlinear analysis varied substantially from the theory for the selection of all wing parameters. How-
exact linear-theory predictions for all combinations ever, comparison of linear-theory estimates with ex-
of geometry and flow parameters under study. For perimental data for wings has shown an inability of
slender wings with highly subsonic leading edges, the the theory to consistently predict the measured re-
nonlinear analysis showed that the zero-lift wave- sults (ref. 1). Linearized theory computational tools
drag correlation relationship was maintained; how-
were limited by the constraints inherent to small
ever, as the wing geometry became nonslender, the
perturbation theory.
flow about the wing became nonlinear and the rela-
An extensive survey of the literature was con-
tionships which define the zero-lift wave-drag corre-
ducted to determine the dominant wing geometric
lation parameter were not maintained. characteristics and flow conditions which should be
The aerodynamic characteristics of delta wings at used in assessing the supersonic aerodynamics of
lifting conditions have been evaluated for the effects wings. The initial result of this effort was the identifi-
of wing leading-edge sweep, leading-edge bluntness, cation of the delta or triangular wing planform as the
and wing thickness and camber and then summarized most likely candidate for future parametric super-
in the form of graphs which may be used to assess the
sonic wing studies due to the extensive experimental
aerodynamics in the preliminary design process. Em-
data base which was available. In addition, the em-
pirical curves have been developed for the lift-curve pirical correlations derived for delta wings could be
slope, nonlinear lift effects, maximum lift, longitu- extended to other simple wing planforms, such as
dinal stability, and distribution of lift between the arrow and diamond wings, through the use of the
upper and lower surfaces of a wing. In addition, the geometric and flow correlation parameters.
impact of various airfoil parameters, wing leading- This paper presents the results of a combined ex-
edge sweep, and lift coefficient on the drag-due-to-lift perimental and theoretical study of the aerodynam-
characteristics has been shown theoretically. ics of delta wings at supersonic speeds along with a
The various graphs which detail the aerodynamic preliminary delta wing selection and design philoso-
performance of delta wings at both zero-lift and phy. The supporting information for this study was
lifting conditions were then employed to define a derived from both the application of a nonlinear in-
preliminary wing design approach in which both the viscid computational method (ref. 2) and from pub-
low-lift and high-lift design criteria were combined to lished force, pressure, and flow visualization data for
define a feasible design space. delta wings.
Results of this study should provide a better un-
Introduction derstanding of the effect that airfoil profile, wing
An extensive survey of the literature reveals that sweep, and Mach number have on the aerodynam-
supersonic aerodynamic research began in the 1940’s. ics of delta wings at supersonic speeds and should
The literature survey showed that during the 1940’s establish the limitations of linear and nonlinear the-
and 1950’s most of the experimental studies at super- ories to predict these effects. The results of the study
sonic speeds were conducted on simple wings, bod- are presented in a parametric fashion in an effort to
i ies, and wing-body geometries as the aerodynami- consolidate the effects of wing sweep, Mach number,
cists attempted to determine the nature of the new and airfoil shape.
flow regime. However, with the introduction of linear
theory in the 1950’s and the emergence of high-speed Symbols
computers in the 19601s, there was a gradual de- A wing aspect ratio
cline in the amount of experimental research directed
toward the study of fundamental flows over simple b wing span
wings and bodies and a large increase in research di-
rected toward the application of linear theory to the CD drag coefficient, %
change in drag coefficient relative to MN component of Mach number normal
flat wing at zero lift to wing leading edge, M cos A
(1 +sin' a tan' A)'/'
drag-due-to-lift parameter
m position of airfoil maximum thick-
zero-lift wave-drag coefficient ness, expressed as a fraction of local
axial integration drag coefficient chord

sectional drag coefficient NRe Reynolds number


9 dynamic pressure
lift coefficient:
R leading-edge radius parameter for
nonlinear incremental change in lift NACA modified four-digit-series
coefficient with respect to linear lift airfoil, Leading-edge radius =
coefficient, C L ~ ~ =-~~OCL,
OO c[l.l019(~R/S)~]
maximum lift coefficient f position of airfoil maximum
lift-curve slope, evaluated at zero thickness
lift, per deg wing reference area
pitching-moment coefficient, local wing cross-sectional area
Pitching moment
nsz wing airfoil thickness
nonlinear increment in pitching- streamwise coordinate
moment coefficient with respect to
streamwise center of pressure
linear pitching-moment coefficient,
dC spanwise coordinate
Cmla=200 - 2 0 e I a = 0 0
normal-force coefficient spanwise center of pressure on the
wing semipanel
wing lower surface normal-force
coefficient YLE spanwise position of leading edge

wing upper surface normal-force 2 vertical coordinate


coefficient a angle of attack
coefficient of pressure ON angle of attack normal to wing
coefficient of pressure at zero lift leading edge, tan-'
minimum wing upper surface P = Jm
pressure coefficient lift-curve-slope parameter, per deg
PCL,
vacuum pressure coefficient, - 2 1 7 ~ ~ p cot A leading-edge sweep parameter
wing chord ratio of specific heats, 1.40 for air
wing mean geometric chord fraction of local wing semispan
wing root chord semispan location of vortex action
marching step size used in nonlinear line
analysis A wing leading-edge sweep angle, deg
longitudinal stability level evaluated 7 airfoil thickness parameter, ex-
at zero lift, per deg pressed as fraction of local chord,
zero-lift wave-drag correlation tl c
parameter, C ~ , w l r ' A Subscripts:
leading edge C cross flow
Mach number ref reference
Discussion Zero-Lift Wave Drag
Supersonic linear theory suggests that the zero-
The present study is directed toward understand- lift wave-drag ( C O , ~of) a delta wing varies as the
ing the fundamental aerodynamic performance of
delta wings through empirical correlations of experi- airfoil thickness-to-chord ratio squared (.r2) and as
mental data and nonlinear computations. Nonlinear the wing aspect ratio ( A ) for a given airfoil pro-
analysis is used only when there are insufficient ex- file. In 1957, Bishop and Cane (ref. 27) incor-
perimental data with respect to a particular geo- porated the dependence of the zero-lift wave drag
metric or flow parameter under investigation. The upon Mach number into the standard linear-theory
data base used in this study is from references 3 drag curves by combining the Mach number param-
through 25 and summarized in table I. eter ,L? with the wing geometric parameter A and
then plotting the linear-theory-dependence parame-
The geometric and flow parameters under in- ter (C,,w/.r2A) against the leading-edge sweep pa-
vestigation are presented first. The zero-lift drag rameter ( p cot A). This method of presentation rep-
results are then presented followed by the lifting resents a plot of zero-lift wave drag against Mach
characteristics. number for a given wing airfoil shape rather than
curves showing the effect on wave drag of one or
more wing geometric parameters as had been done
Aerodynamic and Geometric Parameters
previously (ref. 28).
The primary intent of this section of the paper
The identification of an adequate correlation pa-
is to establish the dependence of the zero-lift wave
rameter is critical to any empirical study. For delta
drag for delta wings on .r2, A , and M for various air-
wings, linear theory indicates that at supersonic
foil profiles. The dependence is established through
speeds all aerodynamic characteristics are a function
nonlinear aerodynamic analysis with the method of
of the leading-edge flow condition, as described by
reference 2 because of a lack of experimental data.
the leading-edge sweep parameter p cot A. In addi-
Nonlinear solutions are only presented for subsonic
tion, the aspect ratio of a delta wing is defined as
leading-edge conditions due to a restriction in the
being four times the cotangent of the wing leading-
present full-potential solution technique. All zero-
edge sweep angle. (See fig. 1.)
lift wave-drag plots are in the format established by
Presented in figure 1 is a plot representing the Bishop and Cane.
range of leading-edge sweep under study, and pre-
sented in figure 2 is a plot of Mach number against Diamond airfoil. Computations of the zero-lift
p cot A for a range of A. Wings with aspect ra- drag of three-dimensional wings with diamond air-
tios varying from a minimum of 0.5 to a maximum foils have dominated the theoretical wave-drag stud-
of 4.0 were selected as being representative super- ies. The simplicity of the geometry lends itself to
sonic planforms (fig. 1). Also indicated in figure 2 a closed-form solution of the linearized equations.
is the sonic leading-edge condition ( p cot A = 1.0) A comparison of the nonlinear and linear solutions
for delta wings. The data of figure 2 show that in- for delta wings with diamond airfoils is presented
creasing aspect ratio significantly reduces the Mach in figure 6. Nonlinear solutions were obtained with
number for a given value of p cot A. the method of reference 2, and the linear solutions
are taken from the closed-form solutions presented
The zero-lift drag study concentrated on the ef- in reference 27. Attempts to exactly duplicate the
fect of airfoil profile on the supersonic aerodynam- closed-form linear-theory solutions of reference 27
ics of delta wings. The study includes a parametric with linear-theory computer codes (ref. 29) were un-
investigation of diamond, circular-arc, and NACA successful because of the inaccuracies inherent to the
modified four-digit-series airfoils (ref. 26) in which numerical techniques employed to solve the equa-
the effect of thickness ratio .r for all airfoils, maximum tions. The best results received from the linear-
thickness position m for the diamond and NACA theory codes had a maximum error, referenced to
modified four-digit-series airfoils, and leading-edge the closed-form solution, of 5 percent. Nonlinear-
bluntness R for the NACA modified four-digit-series and linear-theory curves are presented for positions
airfoil is presented. The range of geometric param- of maximum airfoil thickness (m)of 0.2, 0.3, 0.4,
eters is graphically presented in figures 3, 4, and 5. and 0.5. The linear-theory curves represent the zero-
Thickness ratio varies from 0.02 to 0.10; maximum lift wave drag of all delta wings. The nonlinear curves
thickness position, from 0.2 to 0.5; and leading-edge were computed for a 4-percent-thick delta wing of
1
- -
I radius parameter, from 0.0 to 8.0 within the present aspect ratio 1.0. The delta wing with aspect ratio
l study. of 1.0 was selected for nonlinear calculations because
I
3
of the large variation in Mach number which could be and the curve for m = 0.3 for values of @ cot A be-
analyzed for values of @ cot A from 0.125 to 1.0. A tween 0.275 and 0.55.
review of figure 2 shows that values of /3 cot A be- The large differences between the linear and non-
tween 0.125 and 1.0 correspond to a Mach number linear predictions do not solely establish the existence
range of 1.12 to 4.12 for a wing with A = 1.0 com- of nonlinear flow over the wings, such as large shocks
pared with a Mach number range of 1.01 to 1.67 for or expansions. A comparison of the two methods
a wing with A = 3.0. shows that the linear method tends to predict a sharp
Linear-theory zero-lift wave-drag predictions pressure gradient due to local surface discontinuities,
(fig. 6) show a large effect of the position of airfoil whereas the nonlinear method predicts a much more
maximum thickness and increasing @ cot A. The re- gradual change in pressure. For p cot A < 0.4, the
sults indicate that, for values of @ cot A less than 0.6, flow calculated with the nonlinear theory is sub-
a forward shift in the position of airfoil maximum critical in the cross-flow plane ( M c < l.OO), and
thickness from 0.5 to 0.2 results in a reduction in the forces predicted by linear and nonlinear theo-
drag of 80 percent. For values of @ cot A greater ries are in close agreement. At p cot A > 0.75, the
than 0.8, a rearward shift in the position of airfoil two solutions diverge and a comparison of the cal-
maximum thickness from 0.2 to 0.5 would also pro- culated flow fields, with the nonlinear code, reveals
duce a reduction in drag of 80 percent. For values the development of supercritical cross-flow conditions
of @ cot A between 0.6 and 0.8, the relationship be- ( M c > 1.00). However, for values of @ cot A be-
tween the four linear-theory curves varies drastically. tween 0.4 and 0.75, the flow about the wing is sub-
In this range of @ cot A, the airfoil maximum thick- critical yet the differences between the linear and
ness lines become supersonic (Mach angle exceeds the nonlinear curves are significant. To provide insight
maximum thickness line sweep angle) for each airfoil into the flow condition in this region, spanwise sur-
at different values of p cot A. When the Mach num- face pressure distributions from the nonlinear anal-
ber normal to the airfoil maximum thickness sweep ysis are shown in figure 7 for a delta wing with
line becomes supersonic, a singularity arises in the a 4-percent-thick diamond airfoil and aspect ratio
linear solution which produces a large drag rise (first of 1.0 at a Mach number of 2.60, @ cot A = 0.6. Pre-
peak). The second peak of each linear-theory curve dicted pressure results are presented for airfoils with
corresponds to the drag rise associated with the oc- m = 0.2 and 0.5 at streamwise positions of 20, 40,
currence of a sonic leading-edge condition; this occurs 60, and 80 percent of the root chord. A comparison
at @ cot A = 1.0 for delta wings. As Mach number of the pressures for the two airfoils shows a smooth
is increased beyond the sonic leading-edge condition, and subcritical character to the flow with larger span-
@ cot A > 1.0, linear theory predicts a gradual reduc- wise pressure gradients occurring for the airfoil with
tion in the drag coefficient for all diamond airfoils. m = 0.2. The forward position of the airfoil max-
imum thickness creates an effectively blunter airfoil
Nonlinear zero-lift wave-drag estimates for the ( m = 0.2); this results in larger compression pres- '
effect of the position of maximum airfoil thickness sures at the leading edge and lower expansion pres-
and Mach number are similar to the linear pre- sures as the flow expands around the airfoil ridge line
dictions; however, the details of the four nonlinear when compared with the airfoil with m = 0.5. The 1
curves are significantly different from their linear- equivalent drag values, at @ cot A = 0.6, which re- '
theory counterparts. Both linear and nonlinear the- sult from an integration of the pressure data are due
ories predict a crossover in the four curves; however, to compensating drag characteristics in the stream-
I
the linear-predicted crossover occurs over a range of wise direction. Over the forward portion of the '
@ cot A from 0.6 to 0.8, and the nonlinear-predicted wings (0 < x/cr < 0.3), the local drag of the air- ~

crossover for all curves occurs at @ cot A = 0.6, indi- foil with m = 0.2 is greater than that of the air-
cating that, at @ cot A = 0.6, the drag of a delta wing foil with m = 0.5; between x/c7 of 0.3 to 0.7, the I
with a diamond airfoil is independent of the position drag is similar; and over the aft portion of the wing I
of airfoil maximum thickness. On either side of this (0.7 < x/cr < l.O), the drag of the airfoil with 1
crossover point in the drag curves the effect of the po- m = 0.5 is greater because of the increased slope I
1
sition of maximum airfoil thickness is similar to that at the trailing edge.
predicted by linear theory. In general, the nonlinear
analysis results show a much smoother variation with Selected nonlinear analysis was performed next
@ cot A (no peaks or valleys) and lower drag than the to establish the dependence of the wave drag upon
linear-theory estimates. Nonlinear theory predicts the thickness-to-chord ratio squared (.r2) and aspect
lower drag for all conditions except for the curve for ratio ( A ) for delta wings with diamond airfoils. Pre-
m = 0.2 for values of @ cot A between 0.2 and 0.7 sented in figure 8 is the predicted nonlinear effect for

4
a thickness variation from 0.02 to 0.10 and an as- Except for the diamond profile, it is probably the
pect ratio variation from 0.5 to 3.0. Results are pre- airfoil most frequently used in fundamental studies
sented as a ratio of the nonlinear-predicted zero-lift of wings. Presented in figure 9 is a comparison of
wave drag for each geometry to a reference nonlinear- the axial distributions of area and drag for a series
predicted zero-lift wave drag which has been scaled of equivalent 4-percent-thick sharp airfoils on a delta
by the ratios of the computed to reference wing wing with A = 1.0 at M = 1.80. The three airfoil
aspect ratios (A/Aref) and thickness-to-chord ratio profiles are a diamond with m = 0.5, a circular arc,
(
squared (r/rref)2) : and a sharp modified four-digit series with m = 0.5.
The diamond and modified four-digit-series airfoils
are presented in this section of the paper to provide
a reference point for discussions on the circular-arc
profile. The axial distributions of area and drag
show that the circular-arc airfoil and the equivalent
Thickness effects, presented on the left of figure 8, sharp modified four-digit-series airfoil (m = 0.5,
were evaluated for a reference delta wing with A = R = 0) are similar, and both have a greater volume
1.0, m = 0.5, and an airfoil with r = 0.04 at p cot A and a different axial drag distribution than those for
= 0.25, 0.5, and 0.75. The nonlinear analysis shows the diamond airfoil. The drag characteristics show
a reduced drag with increasing thickness compared that the circular-arc airfoil has higher drag than the
with the value estimated with the zero-lift wave- diamond airfoil at the apex and trailing edge; this is
drag correlation parameter ( K ) . The data on the due to the increased surface slopes in these regions.
left of figure 8 show that, at p cot A = 0.25, the The effect of Mach number on the nonlinear-
linear-theory-dependence relationship is maintained predicted drag characteristics of the circular-arc air-
within 2 percent; for p cot A = 0.5, a 10-percent foil is presented in figure 10. Also shown in the figure
variation is observed; and for p cot A = 0.75, a are the nonlinear drag characteristics of the equiva-
20-percent variation is found. At /3 cot A = 0.75, the lent modified four-digit-series airfoil and the diamond
predicted cross-flow Mach number is supercritical’for airfoil, along with the linear-theory solution for the
the 4-percent-thick wing (figs. 7 and 8); an increase diamond airfoil with m = 0.5. Nonlinear analysis
in thickness would only tend to magnify these effects. shows a smooth variation in drag for p cot A be-
Note that solutions could not be attained for wings tween 0.125 and 1.0. Comparison of the drag levels
with zero thickness (T = 0.00) with the selected shows that the diamond airfoil has the lowest drag,
nonlinear methodology; however, by definition an the modified four-digit-series airfoil has the highest,
exact agreement between the linear and nonlinear and the circular-arc profile has drag slightly less than
solutions would occur. Presented on the right of that for the modified four-digit-series profile. These
figure 8 is the effect of aspect ratio for a reference results are consistent with the data presented in
delta wing with A = 1.0, r = 0.04, and m = 0.2 figure 9.
at p cot A = 0.5 and 0.75. Nonlinear analysis shows Predicted nonlinear effects of airfoil thickness and
a nonlinearly increasing drag with increasing aspect wing aspect ratio for the circular-arc airfoil (fig. 11)
ratio and a maximum perturbation of 5 percent for are similar to those observed for the diamond air-
p cot A = 0.75 and 10 percent for p cot A = 0.5. foil. The analysis shows a decreasing drag value with
Nonlinear analysis of delta wings with diamond increasing airfoil thickness and an increasing drag
airfoils has shown that for p cot A less than 0.5 value with increasing wing aspect ratio. The data
(subcritical range), the zero-lift wave drag varies with also show a greater percentage change in drag for the
the thickness squared and aspect ratio, but for higher circular-arc airfoil than for the diamond airfoil (fig. 8)
values of p cot A (supercritical range), the nonlinear due to increasing /3 cot A. This increased nonlinear
results diverge significantly from the linear-theory- variation in the drag is due to an increased amount of
dependence relationship. Within this supercritical nonlinear-predicted flow over the wing with increas-
range, the nonlinearity of the flow over the wing ing Mach number. The analysis results presented in
increases significantly and is shown to be dependent figures 9, 10, and 11indicate that a different complex
upon both the wing and airfoil geometries. and nonlinear flow structure exists for the circular-
arc airfoil than for the diamond airfoil. In addition,
Circular-arc airfoil. The circular-arc profile is the characteristics for the circular-arc airfoil were
a unique class of airfoils which is often used in the shown to be very similar to those for the equivalent
design of supersonic vehicles. The uniqueness of modified four-digit-series airfoil; this suggests that
the circular-arc airfoil lies in the fact that it has the circular-arc profile can be treated as a subset
only a single design parameter-airfoil thickness. of the modified four-digit-series airfoil family. The

5
~~

analysis clearly shows that the zero-lift drag of sharp area and computed nonlinear drag. Results are pre-
airfoils with the same values of T and m is not equal; sented for a 4-percent-thick airfoil with m = 0.5 at
the circular arc has 15 to 40 percent higher wave drag R = 0, 4.0, and 8.0, which correspond to a leading-
than the diamond airfoil but only 4 percent less drag edge radius, expressed as a fraction of the chord,
than the equivalent modified four-digit-series airfoil. of approximately 0.0, 0.001, and 0.003, respectively.
The area and drag data clearly show that leading-
Modified four-digit-series airfoil. The NACA edge bluntness only affects the forward half of the
modified four-digit-series airfoil was selected for the wing. Drag data show that increasing airfoil blunt-
study of blunt leading-edge airfoils because of the ness produces a localized increase in drag at the wing
flexibility in defining and altering the airfoil pro- apex followed immediately by a rapid reduction in
file. The NACA modified four-digit-series airfoil drag. At x / c r = 0.4, the data show a merging of the
family is defined analytically with the following drag data for the three airfoils which then remain
three parameters: airfoil thickness (T), position of coincident over the remainder of the geometry. The
airfoil maximum thickness ( m ) , and leading-edge crossover in the drag data of the three airfoils which
radius parameter ( R ) .The leading-edge radius of the occurs at the apex of the wing produces a canceling
modified four-digit series is a function of the leading- effect that results in a drag value of approximately
edge radius parameter ( R ) and the airfoil thickness 21 counts (0.0021) for the three airfoils. This inde-
parameter (T) through the relationship: pendent nature of the drag due to changes in leading-
edge bluntness is discussed further later.
Leading-edge radius = c [ l . l 0 1 9 ( ~ R / 6 ) ~ ] The fundamental nonlinear characteristics for the
modified four-digit-series airfoil are established for
This airfoil family is used to evaluate the effect of 4-percent-thick sharp airfoils on a wing with A = 1.0.
leading-edge bluntness in addition to the effects of The dependence of the drag on the thickness, as-
T, m, A , and M on the zero-lift wave drag. Re- pect ratio, and leading-edge bluntness is established
sults of the nonlinear analysis for the diamond and separately.
circular-arc airfoils are referenced. To provide insight Presented in figure 13 are the nonlinear-theory
into the geometric character of this class of airfoils, predictions of drag characteristics for sharp modified
the axial distribution of area and drag for various four-digit-series airfoils with maximum thickness lo-
values of the parameters m and R is presented in cations of 0.2, 0.3, 0.4, and 0.5. Results are pre-
figure 12. Nonlinear results are presented for a 4- sented for values of p cot A between 0.125 and 1.0.
percent-thick delta wing with A = 1.0 at M = 1.41, The nonlinear analysis shows trends similar to those
p cot A = 0.25. The effect of the position of airfoil predicted for the diamond airfoil: smooth variations
maximum thickness is shown at the left of the fig- with ,tl cot A and a crossover in the drag of all ge-
ure for sharp airfoils. The area distribution shows ometries over a range of /3 cot A from 0.65 to 0.8.
that moving the airfoil maximum thickness forward The crossover for the modified four-digit-series air-
results in a forward shift in wing volume and a de- foil occurs at a value of p cot A of approximately 0.7
crease in the maximum cross-sectional area. These compared with a value of 0.6 for the diamond air-
two effects combine to produce a smoother area dis- foil. The data show that, for values of p cot A less
tribution and lower drag. than 0.65, moving the maximum airfoil thickness po-
A review of the axial distribution of drag for the sition from 0.5 to 0.2 reduces the drag by a maximum
airfoil with m = 0.2 shows longitudinal symmetry, of 50 percent, and for values greater than 0.8, a rear-
whereas the drag of the airfoil with m = 0.5 is ward shift in the maximum airfoil thickness position
mostly generated over the final 20 percent of the from 0.2 to 0.5 produces a maximum drag reduction
wing length. The integrated drag values (ref. 2), of 30 percent.
which are listed in figure 12, show that the airfoil The zero-lift wave-drag correlation parameter is
with m = 0.5 has 60 percent higher drag than the evaluated for the modified four-digit-series airfoil in
airfoil with m = 0.2. As Mach number is increased, figure 14. The trends and levels for both the effects
the character of the drag for each airfoil changes of thickness and aspect ratio are similar to the results
dramatically as the compressive pressures begin to for the circular-arc airfoil. The nonlinear predictions
dominate the flow. The apex drag of the airfoil with show a maximum variation of 30 percent for thickness
m = 0.2 would be expected to increase significantly variations from 0.02 to 0.10 and a 15-percent varia-
compared with the airfoil with m = 0.5, and the tion for variations in aspect ratio from 0.5 to 3.0.
trailing-edge drag of both airfoils would reduce. A comparison of these results with those for the
Shown on the right of figure 12 is the effect of circular-arc and diamond airfoils shows an increase in
leading-edge bluntness on the axial distribution of the nonlinearity of the characteristics with increasing

6
findings also support the data presented in figure 15
in which the axial distributions of drag for a wing
with a sharp and a blunt airfoil were shown to cross
over just aft of the wing apex producing a canceling
effect of the local drag.
Comments on zero-lift wave drag. Predicted
nonlinear zero-lift wave-drag characteristics of delta
wings with diamond, circular-arc, and NACA modi-
fied four-digit-series airfoils have been shown to vary
substantially from the characteristics predicted by
sonic wing design studies which were based upon lin- linear theory. The nonlinear analysis suggests that
ear theory would frequently employ a sharp airfoil for slender delta wings at small values of p cot A the
zero-lift wave-drag correlation relationship is main-
tained to a higher degree; however, as the wing as-
pect ratio or airfoil thickness ratio increases, the
flow about the wing becomes more nonlinear and
the zero-lift wave-drag correlation parameter is not
maintained. The delta wing drag curves presented in
figures 6, 10, and 13 should be adequate for the pre-
liminary design studies of delta wings (aspect ratio,
Mach number, and airfoil selection).
Lifting Characteristics
The ability to take into account the effect of airfoil
profile, aspect ratio, and Mach number on the lifting
characteristics of delta wings is extremely important
in selecting the proper wing geometry. The predicted
flow about a delta wing at the zero-lift condition
has been shown to be a strong function of these
parameters, and as the wing is taken to a lifting
condition, the nonlinear effects would be expected to
increase significantly (ref. 30). The existing design
philosophy for delta wings, which is based upon
linear theory, assumes that the lift and drag due to
lift of a flat wing are only a function of the leading-
edge sweep angle and Mach number. The lifting
characteristics are presented in the form of summary
graphs in which the drag, lift, and pitching-moment
characteristics are highlighted.
Lee-side frow characteristics. At zero angle of
attack, the flow over the upper and lower surfaces
of the wing is characterized as attached; however,
as the wing is taken to angle of attack, the lee-side
flow characteristics can change dramatically (ref. 21).
In general the lee-side flow over a wing at angle of
attack can be divided into two categories: attached
or separated. The development and character of
each flow type are dependent upon wing planform,
airfoil profile, and wing camber in addition to the
free-stream flow conditions.
The values of MN and CYN conducive to the ex-
istence of an attached flow over the lee side of a
zero-thick delta wing are shown at the top of fig-
ure 17 (ref. 21). At zero- to low-lift conditions, the

7
flow is subcritical in the cross-flow plane (small cross- vortex normal-force vector should be placed to give
flow velocity) with the primary flow direction being the same wing bending moment as produced by the
streamwise. As lift is increased, the flow turning an- vortex pressures. The data show that the vortex po-
gle about the leading edge becomes greater, thus re- sition depends mainly on angle of attack and leading-
sulting in an acceleration of the flow and an increase edge sweep angle. Typically the vortex moves from
in the inboard flow component. The cross flow will a location near 80 percent semispan to a location
eventually turn streamwise as the flow recompresses. near 50 percent semispan as angle of attack changes
Further increases in lift increase both the cross-flow from 4’ to 20’.
component of the flow and the resultant recompres- The lee-side flow descriptions presented are typ-
sion as shown in figure 17 and will eventually pro- ical for any wing geometry at both subsonic and
duce a cross-flow shock. The occurrence of a cross- supersonic speeds. However, the magnitudes of
flow shock system is an indication of the existence of these individual effects differ substantially with Mach
nonlinear supercritical-type flow. Further increases number. The most notable of these differences are
in lift result in shock-induced separation of the observed in the static pressures measured on the
boundary layer and the formation of a leading-edge wing upper surface. As Mach number increases be-
bubble. yond 1.0, the free-stream static pressure decreases
The second type of flow which occurs on the lee rapidly to significantly reduce the aerodynamic im-
side of a wing is a leading-edge separation system, pact of the expansion over the wing lee side.
which is characterized by a viscous, rotational mass
of air that resides inboard of the wing leading edge. Flat delta wings. This section of the paper draws
(See the lower portion of fig. 17.) At low-lift con- upon several previous empirical correlations (refs. 3,
ditions, a leading-edge bubble develops, and as the 4, 32, and 33) as well as the extensive data base in
angle of attack is increased, the bubble lifts off the order to investigate how wing leading-edge sweep,
wing surface and evolves into a vortex. Further in- Mach number, and angle of attack affect the lift-
creases in angle of attack result in the development ing characteristics of delta wings. The supersonic
of a secondary vortex and eventually a vortex with aerodynamic characteristics of flat uncambered delta
shock occurs. For a vortex system, the vortex body wings are presented in figures 19 through 37. All
is connected to the wing surface via the vortex stem aerodynamic characteristics are presented in a para-
or feed sheet. The feed sheet is also a viscous flow metric fashion in order to consolidate a wide range
region that emanates from the wing boundary layer of geometric and flow conditions and in an effort to
near the wing leading edge. The shape and position provide a set of empirical aerodynamic curves which
of the viscous vortex system are dependent upon the may be used in preliminary supersonic wing design.
flow field external to the vortex which interacts with The characteristic lift-curve slope, nonlinear lift
the vortex system until an equilibrium condition is effects, and maximum lift for flat delta wings are
established. In the steady-state model, the vortex presented in figures 19, 20, 21, and 22. The data
system acts as a physical boundary to the external presented in figure 19 are an extension to the analy-
flow field system boundary. The free-stream flow ex- sis of reference 4 in which additional data sets have
pands around the wing leading edge and follows the been evaluated to establish definitive aerodynamic
contour of the vortex system undergoing an expan- characteristics of delta wings. The experimental
sion followed by a recompression as the flow turns data shown in figure 19 represent a wide range of
about the vortex. wing leading-edge sweep, airfoil thickness, and airfoil
This flow field external to, yet influenced by shape; however, the primary distinguishing feature
the viscous vortex system, is termed the “induced of the data is that of airfoil bluntness, as noted by
flow field.” The extent of the induced flow field is the open and solid symbols. A comparison of the
characterized by a stagnation point or reattachment experimental data with linear-theory results shows
line on the wing upper surface inboard of the vortex
body. Inboard of this induced flow field is the
potential flow field, where the flow is attached in a
excellent agreement; for values of p cot A below 0.5
and between 0.5 and 1.0, the experimental data break
away from the linear-theory curve and show lift- 1

streamwise direction. curve-slope values below those predicted by linear
The effects of angle of attack, Mach number, and theory. The data of figure 20 show that not until a
leading-edge sweep on the vortex location were de- value of /3 cot A of approximately 2.0 do the exper-
tailed in reference 31. (See fig. 18.) As shown in imental data reach the linear-theory level. A close
the sketch, the vortex action line is identified as examination of the data also shows a division of the
the fraction of the local wing semispan qV. The two experimental data sets with the sharp leading-
vortex action line is the position at which the edge data falling below the blunt leading-edge data.

a
The scattering of the two data sets could be due to The characteristics of lift-curve slope presented ir
an effective bluntness that may occur for the sharp- figure 19 were established by evaluating the experi-
edge wings which have large thickness or those which mental data at zero lift. To establish the extent of
have the maximum thickness position near the lead- these linear lift characteristics, data were evaluated
ing edge. The change in characteristic lift-curve slope at an angle of attack of 20" and compared with the
with leading-edge bluntness and increasing p cot A lift coefficients which result from an extrapolation of
indicates a change in the local wing flow character- the experimentally determined lift-curve slope at an
istics from the linear-theory model. Linear theory angle of attack of 20". These results are presented in
assumes that all flow disturbances are weak and are figure 21 as an increment in nonlinear lift as a func-
propagated along Mach lines. For small values of tion of the parameter 4p cot2A. The data clearly
p cot A, the disturbances are weak and the Mach show that only for combinations of low Mach num-
cone is an excellent approximation of the bow shock, ber and high leading-edge sweep (low aspect ratio)
but as the free-stream Mach number is increased (in- does nonlinear increasing lift occur (positive values
creasing /3 cot A), the differences between the linear- of AC,).
theory Mach cone and the finite-strength bow shock This observation was initially reported by Brown
increase; this indicates the growth of a strong non- and Michael in 1954 (ref. 34), Squire, Jones, and
linear disturbance field. This effect tends to produce Stanbrook in 1963 (ref. lo), and Squire in 1967
a change in flow angularity, a loss in energy, and (ref. 35) and 1980 (ref. 36). However, with this
a reduction in the Mach number at the wing lead- present analysis, the bounds of this unique flow
ing edge. Further increases in Mach number, angle condition have been established. For values of the
of attack, or p cot A produce greater changes in all correlation parameter 4p cot2 A between 0.5 and 1.0,
flow conditions and a reduced lift effectiveness com- nonlinear lift effects are not present; for values of
pared with those for the linear-theory model. The in- 4p cot2 A greater than 1.0, nonlinear decreasing lift
creased lift-curve slope with increased wing leading- occurs; and for values of 4p cot2 A less than 0.5,
edge bluntness is attributed to both the increased nonlinear increasing lift is found.
expansion of the flow about the wing leading edge The maximum lift coefficient for thin delta wings
onto the wing upper surface and the increased lower plotted against the parameter M I A is presented in
surface compression pressures. figure 22. The supersonic aerodynamics of delta
Presented in figure 20 are lift-curve slopes pre- wings at high angles of attack is presented to ensure
dicted by nonlinear theory for delta wings with sharp a complete review of all pertinent aerodynamic char-
and blunt leading edges. A comparison of the lin- acteristics. Possible applications of the data at these
ear, nonlinear, and experimental data shows that the characteristics would be to canards or horizontal tails
linear-theory analysis is in very good agreement with of aircraft and fins of missiles. An interesting point to
the experimental data and nonlinear-theory analy- note is that, despite the large range in C L ,which ~ ~
I sis for values of p cot A less than 0.5. For values of was observed in the data (1.1 2 CL,max 2 0.7),
p cot A greater than 0.5, the nonlinear analysis and all data showed that the maximum lift occurred at
experimental data show a lower lift-curve slope than an angle of attack between 40" and 50"; this indi-
that predicted by linear theory. At these large values cates that C L , is~dominated
~ by the rotation of
of the leading-edge sweep parameter, the nonlinear- normal-force vector and not a function of flow con-
theory analysis of wave drag at zero lift (fig. 6) also ditions on the wing upper surface as is true at sub-
shows an increase in the nonlinear characteristics of sonic speeds. Experimental data at subsonic speeds
the flow and a diverging of the two solutions. show that wing upper surface flow separation or vor-
tex breakdown has a large influence on C L , ~The ~ .
The effect of airfoil bluntness on the lift charac- data show that C L ,varies ~ ~from a maximum of 1.1
teristics was studied on the modified four-digit-series at a value of MIA of 0.05 to a minimum value of 0.7
airfoil. Analysis was performed over a range of an- at a value of MIA of 9.0; this indicates the effect
gle of attack from 0" to 10" in which the linearity of both Mach number and wing sweep. The com-
of the lift-curve slope was established within 3 per- bination of high supersonic Mach number and high
cent. The nonlinear analysis and experimental data wing sweep which is required to achieve large values
are in excellent agreement and both compare well of the correlation parameter ( M I A ) corresponds to
with linear theory for values of p cot A less than 0.7. wing flow conditions which would produce reduced
(See fig. 20.) For values of p cot A greater than 0.7, values of C L , ~As~ an . example, experimental data
leading-edge bluntness results in a lift-curve slope be- show that the contribution to the total lift coefficient
low that predicted by linear theory but greater than from the wing windward side is only slightly depen-
that for sharp airfoils. dent of wing sweep and Mach number at maximum

9
lift coefficient. However, the leeward-side contribu- show a variation in both the spanwise and stream-
tion to the lift coefficient is greatly reduced with in- wise center of pressure with changes in wing aspect
creases in both wing sweep and Mach number; these ratio, whereas the data at an angle of attack of 50'
effects combine to produce the characteristics shown (fig. 25(b)) collapse into a single curve. The variation
in figure 22. These individual wing upper and lower in wing center of pressure can be directly related to
surface lift characteristics are discussed in more de- the flow conditions about the wing. At low angles of
tail in the discussion of figures 27 through 38. attack, the lee-side flow characteristics for each wing
vary considerably between attached and separated
The longitudinal stability characteristics of flow; however, as the wings are taken to a = 50°,
uncambered delta wings, presented in figures 23 the flow becomes quite similar and is characterized
through 26, were determined by computing the slope by two wing leading-edge vortices.
across the zero-lift condition. The data have all been The lift and pitching-moment data presented
reduced about the two-thirds root-chord location, the have been shown to be dominated by wing leading-
linear-theory-predicted moment center. Presented edge sweep effects. A review of the existing data
in figure 23 is the longitudinal stability parameter base failed to uncover an adequate parametric set of
P ( d C m / d C L ) evaluated at zero lift plotted against wing data which could be used to isolate the effects
the leading-edge sweep parameter p cot A. Experi- of geometry and flow conditions on the drag-due-to-
mental data are presented for both s h a r p and blunt- lift characteristic; therefore, the nonlinear computa-
leading-edge wings with various leading-edge sweep tional method of reference 2 was employed for this
angles and airfoil shapes over a wide range of Mach analysis.
number. An examination of the data in figure 23 Nonlinear-predicted drag-due-to-lift characteris-
shows a gradual forward movement in the wing cen- tics for blunt airfoils are presented in figures 26
ter of pressure when the wing leading-edge flow con- and 27. Also noted in each figure are the linear-
dition changes from subsonic to supersonic. For the theory zero-thrust and full-thrust boundaries. The
subsonic leading-edge condition, all data show either nonlinear-theory method selected for this study is
neutral stability or positive stability, whereas for the valid only for conditions of attached flow; as a re-
supersonic leading-edge condition, all data show a sult, nonlinear analysis was limited to blunt airfoils at
negative stability level. As mentioned previously, conditions of low-to-moderate lift coefficients (CL 5
these data were obtained at the zero-lift condition. 0.3). The effect of airfoil thickness, airfoil maximum
To evaluate the extent of the linearity of these char- thickness position, and airfoil bluntness on the drag-
acteristics, the zero-lift d C m / d C L level was extrap due-to-lift parameter has been studied for a delta
olated to an angle of attack of 20' and compared wing with aspect ratio 1.0 at a Mach number of 1.41
with the measured experimental data at that angle (fig. 26). Results of the analysis show that increasing
of attack. Presented in figure 24 are the nonlinear thickness, moving the maximum thickness location
increments in pitching moment evaluated at an an- forward, or increasing airfoil bluntness improves the
gle of attack of 20'. The data show a trend with lifting efficiency of the airfoil, with airfoil thickness
increasing p cot A which is opposite that for the sta- providing the largest improvement and bluntness the
bility data of figure 23. For P cot A less than 1.0, smallest. Comparison of the results shows that the
the data show a positive increment in pitching mo- nonlinear theory predicts a lower drag-due-to-lift pa-
ment, and for p cot A greater than 1.0, the data show rameter value than that predicted by linear theory for
negligible increments in pitching moment. The data uncambered wings (no thrust), and for low values of
of figure 23 indicate that, for the subsonic leading- the lift coefficient (CL < O . l ) , the nonlinear theory
edge condition at zero lift, the majority of the data predicts drag characteristics below the linear-theory
show that the wing stability level was approximated optimum (full thrust). The increase in the drag-due-
well by linear theory; however, as the wing is taken to-lift parameter with increasing lift coefficient is a
to an angle of attack, the data of figure 24 show result of a loss in the aerodynamic thrust force as the
the development of a nonlinear increment in pitch- wing rotates through a range of angles of attack. At
ing moment for the subsonic leading-edge condition, low-lift conditions, the flow about the wing surface
and this suggests the development of nonlinear flow is characterized by a gradual expansion about the
over the wing. nose of the airfoil followed by a smooth recompres-
The longitudinal stability characteristics are pre- sion on the upper surface; this produces significant
sented in figures 25(a) and 25(b) where the stream- amounts of aerodynamic thrust. As angle of attack is
wise and spanwise wing centers of pressure on a increased, the local expansion about the airfoil nose
single wing panel are presented for angles of attack increases significantly and extends over a greater por-
of 5' and 50°, respectively. The data of figure 25(a) tion of the airfoil leeward surface: this results in a

10
smaller percentage of the local expansion pressures leading-edge pressure reduces slightly to increase the
acting on the nose of the airfoil. This expansion re- aerodynamic thrust force. In addition, the thrust
sults in a reduction in the aerodynamic thrust force of force which is produced normal to the wing leading
the wing. The favorable flow conditions which exist edge varies as the cosine of the leading-edge sweep
about the nose of a blunt airfoil (low pressures) must angle. As aspect ratio is increased the wing sweep
coincide with favorable flow conditions over the aft decreases to increase the aerodynamic thrust force.
section of the airfoil (high pressures) if an improve- In general, the character of all the curves presented
ment in the drag-due-to-lift characteristics is to be in the figure is a rapid reduction in the drag-due-to-
realized. lift parameter (increased performance) with increas-
The improvement in drag-due-to-lift characteris- ing p cot A up to a value of approximately 0.6. For
tics with increasing airfoil bluntness is contradictory values of p cot A greater than 0.6, the curves tend
to the results of previous studies (ref. 37) in which to flatten out. This characteristic seems to be quite
complex wing geometries were employed in the study sensitive to aspect ratio and lift coefficient; however,
of the leading-edge thrust phenomenon. However, as a value of p cot A of 0.6 seems to be a representative
discussed, the variation in drag due to lift is not solely mean value for the analysis conducted.
dependent upon the loading at the leading edge of the The previous analysis reviewed the aerodynamic
wing, but it is dependent upon the total wing load- forces and moments of flat delta wings at lifting con-
ing. The nonlinear analysis presented in figure 26 ditions in which the lift and pitching-moment char-
was performed at conditions which were previously acteristics were shown to be dominated by the flow
determined to have a subcritical-type flow structure condition at the wing leading edge ( p cot A) and the
over the wing ( p cot A = 0.25); however, the drag- drag to be equally sensitive to all geometric and flow
due-to-lift analysis showed a highly nonlinear char- parameters. To provide further insight into these ob-
acteristic. To investigate these effects further, non- served characteristics for flat delta wings, wing upper
linear analysis was performed on both a delta wing and lower surface static-pressure distributions are re-
with A = 1.0 and one with A = 2.0 with blunt mod- viewed. The experimental pressure data are reviewed
ified four-digit-series airfoils (T = 0.40, m = 0.5, and to investigate the effects of wing leading-edge sweep,
R = 4.0). Analysis was conducted for values of leading-edge bluntness, wing thickness, wing camber,
p cot A from 0.25 to 0.8. Results of this analysis are Mach number, and angle of attack on the total wing
presented in figure 27 in which curves of constant loading and the individual upper surface and lower
lift coefficient for both wings are shown. Also pre- surface wing loadings. This study makes use of the
sented in the figure are the linear-theory zero-thrust known existence of conical flow for conical geometries
and linear-theory full-thrust curves of the drag-due- at supersonic speeds to extend the use of the limited
to-lift parameter. The nonlinear analyses show an in- amount of wing surface pressures in an effort to r e p
crease in the drag-due-to-lift parameter with increas- resent the total wing loading. However, before this
ing lift coefficient, decreasing p cot A, and decreasing analysis can be generally applied, the lower Mach
aspect ratio. Results for the wing with A = 1.0 at number bounds of the supersonic flow regime need
CL = 0.1 show a variation in the aerodynamic thrust to be identified in order to determine where the con-
from 70 percent of the full-thrust linear-theory value ical flow assumption may be used. In order to deter-
at /3 cot A = 0.25 to 0 percent at p cot A = 0.75. The mine the lower Mach number bound of the supersonic
drag-due-to-lift parameter for the wing with A = 2.0 flow regime, subsonic and transonic data need to be
at CL = 0.1 varied from 114 percent to 100 per- reviewed to ensure that the characteristics of these
cent at p cot A = 0.25 to 0.75. Similar character- two flow fields are differentiated from those at su-
istics to those observed at CL = 0.1 were also ob- personic speeds. Presented in figure 28 are spanwise
served for the two wings at CL = 0.2. An increase pressure distributions at various longitudinal ( x / c r )
in drag due to lift with decreasing wing aspect ra- stations for a delta wing with A = 1.0 at Mach num-
tio would normally be expected for a constant Mach bers from 0.60 to 3.50. The pressure data clearly
number; however, by presenting the data with re- show that nonconical flow conditions exist at both
spect to p cot A, wing leading-edge sweep or aspect M = 0.60 and 0.90. However, between M = 0.90
ratio effects should be factored out of the solution. and 1.20, the pressure data change from a nonconical
As a result, the analyses suggest that Mach number nature to a very nearly conical flow on both the up-
has a large impact on the aerodynamic performance per and lower surfaces. At a Mach number of 1.60,
of a given wing. As Mach number is reduced, the the spanwise pressure distributions are conical and
characteristics around the wing change dramatically; remain conical for all further increases in Mach num-
the upper surface leading-edge suction pressure coef- ber. Based upon these data, a minimum supersonic
ficients become more negative and the lower surface Mach number is defined as approximately 1.20. An

11
~~

interesting point to note is the dramatic change in supersonic speeds is a combination of nonlinearly de-
the upper surface pressures and the comparative in- creasing upper surface normal force and nonlinearly
sensitive nature of the lower surface pressures with increasing lower surface normal force. These results
increasing supersonic Mach number. These effects suggest that only at subsonic speeds ( M 5 0.90) can
are discussed further in the discussion of figure 29. the lift increment between the linear potential-theory
solution and experimental data belong solely to up-
Typical effects of Mach number and angle of per surface vortex-induced effects. At both transonic
attack on the spanwise pressure distribution for a and supersonic speeds, the lift increment is probably
delta wing with A = 1.0 are presented in figures 29 due to a combination of both nonlinear lower surface
and 30, respectively. To assist in the discussion of and nonlinear upper surface effects. In particular, at
the pressure data only a single pressure distribution supersonic speeds, the upper surface vortex-induced
is presented in figures 29 and 30. In addition, the lift increment reduces with increasing angle of attack,
data of figures 29 and 30 are used to represent the and the lower surface compression lift increment in-
total wing pressure distribution (conical flow). creases with increasing angle of attack.
If the assumption is made that conical flow exists, At both subsonic and supersonic speeds, the lift-
then the data presented in figure 29 show that, at a curve slope and lifting efficiency of a wing are directly
constant angle of attack of 15', increasing the Mach related to the aspect ratio of the wing. To determine
number from 0.6 results in an ever increasing reduc- the effect of aspect ratio on the distribution of lift
tion in upper surface normal-force coefficient. A re- between the upper and lower surface of a wing, the
versal of these upper surface effects is found on the analysis presented in figures 30 and 31 was repeated
lower surface where the lower surface normal-force for a wing with A = 0.5 (A = 82.87') and one
coefficient increases with an increase in Mach num- with A = 2.0 (A = 63.43'). Presented in figure 32
ber as the transition is made to supersonic speeds are spanwise surface pressure distributions for wings
and then remains constant. As Mach number in- with A = 0.5, 1.0, and 2.0 at angles of attack of
creases, the increase in lower surface lift coefficient 10' and 20' for M = 0.60 and 1.60. Comparison
does not completely compensate for the reduction in of the upper and lower surface pressure plots in
upper surface lift, and the result is a reduction in to- figure 32 shows that the supersonic data vary in a
tal wing lift coefficient with increasing Mach number more orderly fashion with increases in aspect ratio.
at a constant angle of attack. The combination of a The data at both M = 0.60 and 1.60 show an increase
reduction in the upper surface lift coefficient with an in lift with increasing aspect ratio, with the upper
increase in lower surface lift coefficient highlights the surface dominating at subsonic speeds, but a more
impact of compressibility and the vacuum pressure equal distribution is evident at supersonic speeds.
coefficient at supersonic speeds. These effects are quantified in figure 33 in which the
Presented in figure 30 is the effect of increasing percentage of lift on the upper and lower surfaces is
angle of attack on the spanwise pressure distribu- depicted. The graph shows that Mach number has
tion at M = 0.60, 0.90, 1.20, and 1.60 for the delta the largest impact on the distribution of lift for all
wing with A = 1.0, The pressure data show that at wings at angles of attack of 10' and 20'. All wings
M = 0.60 and 0.90, both the upper and lower surface show a shifting of lift from the upper surface to the
lift increase proportionally. The data for M = 1.20 lower surface with increasing angle of attack except
and for M = 1.60 indicate a shift in the percent- for the wing with A = 0.5 at M = 1.60, which shows
age of lift force from the upper surface to the lower a reverse effect. The combination of low aspect ratio
surface. This effect is quantified in figure 31 which and low free-stream Mach number creates conditions
depicts the percentage of lift on both the upper and favorable to the development of nonlinear lift effects.
lower surfaces for angles of attack of 10' and 20'. (See fig. 21.) These conditions combine to give
The graph of figure 31 shows that at subsonic speeds a value of the nonlinear lift correlation parameter
( M = 0.60 and 0.90), approximately 70 percent of (4p cot' A) of 0.08, which is a condition at which
the total lift comes from the upper surface, indepen- large amounts of nonlinear lift would occur. Based
dent of angle of attack. At M = 1.20 and Q = lo', upon the data in figure 33 it may be concluded that
the upper surface carries 70 percent of the lift, but this increment can be attributed to the wing upper
at (Y = 20°, this is reduced to 60 percent. This trend surface flow field. A comparison of the wings with
is also observed at M = 1.60 where at Q = 10' the A = 1.0 and 2.0 at M = 1.60 shows a gradual
upper surface dominates with 56 percent of the lift increase in lower surface dominance with increasing
and at Q = 20' it has been reduced to 48 percent wing aspect ratio at both angles of attack. The
of the lift. The shifting of lift from the upper to pressure data of figure 32 show that this transition in
the lower surface with increasing angle of attack at lift force is primarily due to an increase in the lower

12
surface contribution to the normal force and not due of attack. The lower surface normal-force coefficient
to upper surface effects. C& is plotted as a function of the normal angle of
A summary of the individual normal-force char- attack CYN and the data collapse into a family of
acteristics of the upper and lower surfaces for thin constant leading-edge sweep curves. Each of these
delta wings is presented in figures 34, 35, and 36. curves is comprised for a range of Mach numbers,
Figures 34 and 36 are formulated by integrating with the only limitation being that the normal Mach
experimental spanwise pressure distributions to ex- number M N of all these data be less than 1.00. How-
tract sectional upper and lower surface normal-force ever, the data have a maximum variation with Mach
coefficients. The sectional normal-force coefficients number of approximately h O . 0 3 C ~ .The data show
are then used to represent the total wing upper or that the lower surface produces a nonlinearly increas-
lower surface normal-force coefficient based upon the ing normal-force increment with increasing angle of
known existence of conical flow for delta wings at su- attack, and it can be seen that the nonlinearity in-
personic speeds. The curves of figures 34, 35, and 36 creases with an increase in leading-edge sweep.
cover a range of Mach number from 1.50 to 3.50 and An evaluation of the data of figures 34 and 36 sup-
leading-edge sweep of 52" to 85". The results could ports the findings previously observed in figures 21
not be produced for Mach numbers below 1.50 be- and 33, which indicate that nonlinear lift is most
cause of insufficient experimental data. As shown in pronounced for extremely highly swept wings at low
figure 34, when the upper surface normal-force coef- Mach numbers. For these very low values of the pa-
ficient C& is plotted as a function of the parameter rameter p cot A, the highly nonlinear character of
CYNPcot A, the data collapse into a family of constant the lower surface normal-force coefficient adds to the
Mach number curves. The large effect of Mach num- linear character of the upper surface normal-force
ber on upper surface normal force is clearly shown; coefficient and produces a total lift force which in-
for example, an increase in Mach number from 1.50 creases in a nonlinear sense. Similarly, figures 34
to 2.00 (fig. 34(b)) reduces the upper surface lift- and 36 can be used to show that the reduction in lift-
ing potential by 50 percent at a given value of the curve slope with an increase in leading-edge sweep is
correlation parameter. The large reduction in 'up- primarily a lower surface-dominated effect and the
per surface lifting capability with increasing Mach increase in lift-curve slope with a decrease in Mach
number is due to the inability to achieve low val- number is an upper surface-dominated effect.
ues of the upper surface suction pressure coefficient.
The characteristics of the data of figure 34 also indi- To further extend these findings, the effects as-
cate that, for a given Mach number and leading-edge sociated with varying leading-edge bluntness and
sweep, an increase in angle of attack results in an Reynolds number ( N R ~on ) the minimum upper sur-
upper surface normal-force coefficient that increases face pressure are presented in figures 37 and 38, re-
nonlinearly with a decreasing slope. spectively. The data for C;,min are plotted as a func-
tion of the parameter CYNPcot A for flat wings at
Presented in figure 35 is the upper surface mini- conditions in which the flow has separated at the
mum pressure coefficient C&in plotted as a function leading edge. A summary of the effect of leading-
of the parameter CYNPcot A. These data also collapse edge bluntness on C;,min is presented in figure 37(a)
into a family of constant Mach number curves. Also in which curves are presented for both sharp and
noted in this figure is the percent of vacuum limit blunt leading-edge wings at Mach numbers of approx-
which was attained for that particular Mach num- imately 1.50, 1.60, 1.90, and 2.40. The blunt leading-
, ber. The data show that the percent of the vacuum edge curves were taken from figures 37(b) and 37(c),
pressure actually attained is reduced with increasing and the sharp leading-edge data curves were taken
Mach number. At a Mach number of 3.50 (fig. 35(a)), from figure 35. The data show that leading-edge
only 75 percent of the vacuum limit pressure co- bluntness increases the minimum upper surface pres-
efficient was reached; however, at a Mach number sure coefficient for a given value of CXNPcot A but
of 1.50, 97 percent of the vacuum limit pressure does not affect the minimum value of C;,,. The in-
coefficient was attained (fig. 35(b)). creased upper surface suction pressures with leading-
The variation in the lower surface normal-force edge bluntness would be expected to increase the lift-
coefficient is presented in figure 36. Unlike the data curve slope and allow for the development of a given
of figures 34 and 35, which showed that the upper level of lift at a lower wing incidence angle, and, thus,
surface characteristics are a function of leading-edge result in lower drag. These results are directly re-
sweep, Mach number, and angle of attack, the lower flected in the data of figure 19 where the lift-curve
surface characteristics were found to be predomi- slope is shown to increase over that for a sharp wing
nantly a function of leading-edge sweep and angle with the addition of leading-edge bluntness.

13
The data of figure 38 show that at a constant The effect of wing thickness on the wing leading-
Mach number increasing N R ~ for a blunt leading- edge vortex position is presented in figure 39. The
edge wing increases the minimum upper surface pres- zero-thickness flat-wing curve shown in the figure has
sure coefficient for a given value of CYN,O cot A; how- been taken from the flat-wing data of reference 31.
ever, these effects decrease with an increase in Mach The data clearly show an outward movement of the
number. vortex with increasing leading-edge slope at all con-
These data show that at supersonic speeds signif- ditions evaluated. The thick-wing data show that
icant amounts of nonlinear flow exist on flat, sharp- the wing leading-edge angle acts to delay the onset
leading-edge delta wings, and that the nonlinear of flow separation to a higher angle of attack. For a
characteristics of delta wings are affected by leading- constant angle of attack, the data show a weaker vor-
edge bluntness and Reynolds number. These non- tex which is located more outboard compared with
linear characteristics which occur on both the upper the flat-wing data (fig. 42). A comparison of the
and lower surfaces of flat wings are summarized in data for the two wings with a 30' leading-edge sweep
the form of graphs to provide a fundamental under- angle shows a slight shift in the vortex position due to
standing of the aerodynamics of delta wings to the thickness. The data indicate that wing leading-edge
designer. angle and not wing thickness is the dominant mecha-
nism which controls vortex formation, strength, and
Thickness effects. The figures discussed in the position. Despite these large effects on the lee-side
previous section, which summarized the global and flow characteristics between the three wings, there
local aerodynamics of flat delta wings, relied heavily were no noticeable changes on the total wing lifting
on the known existence of conical flow for their de- characteristics. (See fig. 40.) For values of j3 cot A
velopment. The figures were constructed from data below 0.5, the lift-curve-slope data show an insensi-
sets obtained with zero-thick delta wings (conical ge- tivity to thickness; for values of /3 cot A above 0.5,
ometries) or very slender geometries which are near the data show an increase over the flat delta data
conical in nature. In order to study similar aerody- but yet show no variation between the three thick
namic effects for typical wing thickness variations, wings. The lift-curve-slope data were evaluated be-
it would be necessary to have an extensive amount tween cx = 0.0' and 2.05' in which the flow is at-
of upper surface and lower surface pressure distri- tached and both surfaces have positive pressure co-
butions in order to resolve the nonconical flow condi- efficients. (See fig. 44.) The lift curve of these thick
tions. However, if the thick-wing geometry is conical, wings was found to be quite different from that of the
the flow may be assumed to be conical and a single flat wings. In particular, the linearity of the lift curve
spanwise pressure distribution is all that is required for all wings was found to be limited to a few degrees
on both the wing upper surface and lower surface to angle of attack. Moreover, the data of figure 41 show
represent the total wing flow field. that the bounds established for the development of
In an attempt to assess the effect of leading-edge nonlinear lift for thin flat wings are not applicable
angle, a comparison is made between the 30' dia- to thick wings because significant amounts of non-
mond wing and the 60' biconvex wing. In an attempt linear lift are evident well within the flat-wing linear
to assess thickness effects, a comparison is made be- lift region. An examination of the surface pressure
tween the 30" biconvex wing and the 30' diamond coefficient data indicated that the increased lift ef-
wing. Presented in figures 39 through 48 are the fect may be attributed to the wing upper surface.
effects of thickness on the total wing aerodynamics At zero lift the thick-wing upper surface has large
and the local wing loadings for several conical ge- positive pressures, and as we have already seen from
ometries. The thick-wing data were obtained from the flat-wing data, there is a minimum upper sur-
reference 23 in which three delta wings with A = 1.0 face pressure which may be attained at a given Mach
differing in thickness and cross-section shape were number. The differences between these two pressure
tested between M = 1.30 and 2.80. As shown in the levels define the maximum attainable upper surface
inset sketch of figure 39, the three different thick- wing loading. If we compare this value to that for
nesses correspond to a diamond cross section with a zero-thick or flat wing, at zero lift the flat wing
a 30' leading-edge half-angle, measured in the cross- has a pressure coefficient value of approximately 0.0
flow plane, and two biconvex cross sections with 30' compared with a value of 0.05 to 0.10 for the thick
and 60' leading-edge half-angles, also measured in wing. (See fig. 44.) This difference would reduce the
the cross-flow plane. It is recognized that these cross allowable change in upper surface pressure coefficient
sections do not form typical airfoil shapes; however, for the flat wing and, thus, the resultant wing lifting
they probably represent reasonable upper bounds on potential. These effects are quantified in figures 42
airfoil leading-edge angles and airfoil thicknesses.

14
I through 44 for changes in thickness, Mach number, constant Mach number in which the data agree
and angle of attack. reasonably well over the range of the correlation
1 The lower surface pressures show an equal effect parameter Q N P cot A. At low values of the correla-
1 of both the wing leading-edge angle and wing thick- tion parameter, the thick-wing results fall below the
ness. Mach number effects are shown in figure 43 for flat-wing curves, and as the correlation parameter is
the 30' biconvex wing geometry at an angle of attack increased the thick-wing data increase more rapidly
of 8'. The data show trends similar to those observed and eventually cross over the flat-wing curves. The
for the flat delta wings, a reduced upper surface load- nonlinearity in the thick-wing data curves can be
ing with increased Mach number, and a general in- attributed to the large positive pressure coefficients
sensitivity to Mach number on the lower surface. A which exist at Q = 0'. Despite these differences be-
comparison of the thick-wing lower surface data with tween the two data sets, some of the general trends
the flat-wing data (fig. 29) shows a reversed variation of the thick wings agree reasonably well with the flat-
\ with increasing Mach number. The flat-wing data in- wing results.
! dicate a slight increase in lower surface pressure with The effect of the pressure at Q = 0' is high-
increasing Mach number whereas the thick-wing data lighted in figure 47, which presents the upper sur-
, show a decrease in pressure. This pressure reduction face minimum pressure coefficient at M = 1.30, 2.00,
t, is actually misleading, because the pressure data at and 2.80. The data show a significant delay in the
an angle of attack of 0' also show a pressure reduc- occurrence of a negative pressure coefficient due to
I
tion with increased Mach number, which occurs at a increased thickness and leading-edge slope. This ef-
I
rate equal to that observed at angle of attack; this fect is most evident for the data at M = 1.30 where
results in an invariant lower surface loading. there is a 0.25 variation in the minimum pressure co-
Presented in figure 44 is the effect of angle of efficient between the three wings at an angle of attack
attack on the spanwise pressures for the 30' biconvex of 12'.
geometry at M = 1.30. The data show trends The effect of thickness on the lower surface
similar to those for the flat wings; however, the levels normal-force coefficient compares well to the flat-
are quite different especially at an angle of attack wing results of figure 36 (fig. 48); this indicates that
of 0'. As mentioned previously, the positive upper lower surface normal-force characteristics are inde-
surface pressure coefficient at Q = 0' for the thick pendent of wing thickness and Mach number.
wings allows for a greater increment in upper surface
'
I
loading with increased angle of attack. In particular,
the data show an average pressure coefficient of 0.075
Cumber eflecrs. In designing a wing for efficient
supersonic flight, the geometric parameter which is
1 at an angle of attack of O', which is approximately usually optimized after the wing planform selection
9 percent of the maximum flat-wing upper surface is the camber and twist distribution. The purpose
lift force, with total vacuum pressure coefficient on of this section is to evaluate the following: the
1 the wing upper surface being assumed.
~ ~~
wing lift-curve slope sensitivity to camber, the effect
, of aspect ratio on nonlinearities in the flow, and
The distribution of loading between the upper the lift distribution between the upper and lower
and lower surfaces of the 30" biconvex wing at Q =8O
surfaces relative to that for a flat wing. These points
and 16' for Mach numbers of 1.30, 2.00, and 2.80 is were addressed by analyzing a parametric set of six
shown in figure 45. The data show a gradual shifting conical, zero-thickness, cambered delta wings tested
in lift from the upper surface to the lower surface at M = 1.90. The geometries consisted of 15'
with increasing Mach number and angle of attack. If streamwise leading-edge deflections of the outboard
the thick-wing results of figure 45 are compared with 10 percent and 20 percent of the semispan for each
the flat-wing data of figures 29 through 31, we see a of the three leading-edge sweep angles of 75', 67.5',
delay of the dominance of the lower surface on the and 58.25' (ref. 24). These wings were used to
total lift with increasing Mach number. evaluate the effect of camber only, since no twist
A summary of the individual normal-force char- was applied to the spanwise sections. The data are
I acteristics of the upper and lower surfaces for the presented in a similar fashion as those for thickness
i
t
three thick delta wings is presented in figures 46,
47, and 48. Figures 46 and 48 were formulated in
effects in figures 49 through 57.
Unlike the thick-wing data which showed a uni-
the same manner as those for the flat wings. Thick- form progression in the location of the vortex with
~
wing upper surface normal-force data are presented all parameters, the cambered wing data were very
in figure 46 along with the flat-wing curves from erratic (fig. 49). A review of the pressure data re-
1 figure 34 for Mach numbers 1.30, 2.00, and 2.80. vealed that the lee-side flow characteristics were quite
The thick-wing data collapse into three groups of complex due to the sudden expansion about camber

15
hinge line. (See figs. 52, 53, and 54.) However, that observed for the thick wings. Also evident are
once flow separation occurs at the wing leading edge the large hinge-line-induced expansion pressures on
(a > lo’), the flow behaves similar to that for the the wing upper surface (a= 8’) and the oscillatory
thin flat wings. The data also show that increasing nature of the lower surface pressure due to the irreg-
wing camber delays the formation of a wing leading- ular geometry. As with the thick wings, an integra-
edge vortex. These results are analogous to the effect tion of the pressures to obtain the upper and lower
of increasing the leading-edge sweep angle for thick surface wing loadings resulted in values comparable
wings. However despite these flow complexities the with those for the flat wings (figs. 55 and 57).
agreement between the flat-wing and cambered-wing The upper surface normal-force data for all cam-
lift-curve slopes was quite surprising (fig. 50). The bered wings are presented in figure 55. The data are
cambered-wing data show a variation in lift-curve shown to collect near or below the flat-wing data at
slope similar to that observed for thick wings with an M = 1.90 from figure 34 (solid line) and show little
increased lift-curve slope for p cot A greater than 0.5. effect of leading-edge sweep or camber. Presented
This increased lifting efficiency at the higher values in figure 56 are the upper surface minimum pressure
of /3 cot A may be attributed to the increased load- coefficients for all cambered wings. The data fall into
ing potential of the wing leading edge. The cambered a broad band which lies below and parallel to the flat-
wing at an angle of attack of 0’ would have a large wing data. This large scatter of the C&in data can
negative load on the deflected portion of the wing be attributed to the interaction between the hinge-
leading edge. The magnitude of this load at a = 0’ line expansion and the leading-edge separation and
would be equivalent to the additional lift increment produces both a wide range and a dramatic change
available to the cambered wing compared with the in the lee-side flow with changes in camber, wing
flat wing. This effect is supported by the compar- sweep, and angle of attack. The effect of camber
ison of flat and cambered nonlinear lift increments on the lower surface normal-force coefficient is pre-
which show the existence of nonlinear lift for all val- sented in figure 57. The lower surface characteristics
ues of the correlation parameter (fig. 51). To further are shown to separate according to the wing leading-
evaluate the local aerodynamics of cambered delta edge sweep as indicated by the flat-wing data (solid
wings, detailed upper and lower surface pressure co- line), and the data show little effect of camber. A
efficient (at s / ~ x
, 0.66) distributions highlighting comparison of the two data sets seems to indicate a
the individual effects of camber, leading-edge sweep, reduction in the lower surface normal force due to
and angle of attack are presented in figures 52, 53, camber; however, the cambered delta wings of refer-
and 54, respectively. ence 24 were constructed by a deflection of the out-
The effect of camber on the spanwise surface board segment of the reference flat wing. This would
pressure distributions is presented in figure 52 for result in leading-edge sweep angles for the cambered
the 75’ swept wing at an angle of attack of 8’. wings which are greater than those listed.
A significant variation in the leading-edge upper
surface pressures is evident due to increasing camber; Comments on lifting characteristics. The aero-
however, the loadings are equivalent. On the lower dynamic characteristics of delta wings at lifting con-
surface the spanwise pressure distributions show less ditions have been evaluated for the effects of wing
sensitivity to camber. leading-edge sweep, leading-edge bluntness, and wing
Presented in figure 53 is the effect of leading-edge thickness and camber and then summarized in the
sweep on the spanwise pressure distributions at an form of graphs which may be used to represent the
angle of attack of 8’. The data show that the major aerodynamic characteristics of delta wings. Empiri-
influence of leading-edge sweep is an increase in the cal correlation curves derived from experimental data
lower surface loading with decreasing sweep; this also have been developed for the lift-curve slopes, nonlin-
correlates well with the flat-wing data of figure 32(b). ear lift effects, maximum lift, longitudinal stability,
There is also a significant influence on the position and distribution of lift between the upper and lower
of the vortex. surfaces of a wing. However, the impact of airfoil
The variation in the spanwise surface pressure thickness, maximum thickness position, leading-edge
distributions with angle of attack for cambered wings bluntness, wing leading-edge sweep, and lift coeffi-
is presented in figure 54. The data clearly show the cient on the drag and the drag-due-to-lift character-
loading at the wing leading edge for an angle of attack istics is shown theoretically.
of O’, which was alluded to earlier. The integrated
lift increment between the upper and lower surface
Real-Flow Wing Design
spanwise pressure distributions at a = 0’ provides Wing design studies at supersonic speeds have
an additional increment in lift to the wing similar to typically been prefaced with the terms “linear” or

16
“nonlinear” depending upon the theoretical method indicate the point of optimum performance. How-
which was employed within the investigation. In ad- ever, the data of figure 27 show that the drag-due-to-
dition, the previous nonlinear studies have employed lift characteristics are much more complex. Whereas
traditional linear-theory rules in planform and airfoil the data of figure 20 showed no distinguishable effects
selection. Previous supersonic wing design studies of aspect ratio, the drag data show large variations
have also typically employed airfoils with thicknesses in performance due to both wing sweep and lift co-
between 3 and 6 percent and the maximum thick- efficient. The data show that the value of p cot A
ness located at 40 percent to 50 percent chord. The at which the drag-due-to-lift characteristics reach a
aerodynamic reasoning for selecting or limiting the minimum varies between 0.4 and 0.8 depending upon
selection of wing geometries to these geometries has lift coefficient and aspect ratio. A review of the zero-
never been documented. In fact, linear-theory aero- lift wave-drag data (fig. 13) shows that a value of
dynamic graphs which are typically used in prelimi- p cot A of 0.4 corresponds to the bucket of the curve
nary design indicate that significant drag reductions for m = 0.2; as a result, the design space is ex-
may be attained with alternate airfoil shapes. (See panded to values of p cot A between 0.4 and 0.8 for
fig. 6.) The intent of this section of the paper is to completeness.
employ the empirically derived graphs of the previ- The typical application of thickness to uncam-
ous sections in a systematic fashion to select wing bered delta wing results in a geometry which is con-
geometric characteristics conducive to high levels of ical about the wing tip. This classical application
aerodynamic performance. of thickness to a swept wing is no doubt less than
The zero-lift, low-lift, and high-lift aerodynamic optimum for supersonic flight. Experimental data
characteristics previously presented for delta wings (ref. 38) and theoretical analysis (ref. 2) show that
show a large influence due to changes in wing ge- the flow over a swept wing at supersonic speeds tends
ometry and flow conditions. At zero lift, theoreti- to be conical about the wing apex and not coni-
cal analysis with an inviscid nonlinear method shows cal about the wing tip as observed in subsonic flow
significant variations from the standard linear-theory for unswept wings. The conical nature of the flow
curves due to changes in airfoil shape, airfoil thick- field over the wing produces favorable and unfavor-
ness, leading-edge bluntness, and wing leading-edge able pressure fields on the wing surface. For a wing
sweep. A review of the zero-lift-curve drag analysis at lift, the flow over the wing upper surface would
shows that for a cruise or low-lift-dominated design be characterized by an expansion over the leading
the diamond airfoil could provide 20-percent to 50- edge which is followed by a recompression to a more
percent reductions in wave drag compared with the positive pressure as the flow moves inboard. The
NACA modified four-digit-series airfoil. However, if location of the recompression has been observed to
the wing is to operate over a range of Mach number lie along a ray emanating from the wing apex, in-
and lift coefficient the problem becomes more com- dependent of the wing geometry. If the upper sur-
plex and a blunt airfoil should be employed to im- face is divided into four quadrants, defined by the
prove the drag-due-to-lift characteristics. The zero- intersection of the airfoil maximum thickness line
lift wave-drag analysis for the NACA modified four- and the recompression line, two favorable and two
digit-series airfoil (fig. 13) shows that for values of unfavorable performance regions may be identified.
p cot A between 0.6 and 0.8 locating the airfoil max- The two unfavorable regions, which contribute to the
imum thickness at the 20-percent chord provides low drag, are the inboard forward portion and the out-
drag, and the analysis presented in figure 26 shows board aft portion of the wing. The inboard forward
an improvement in the drag-due-to-lift characteris- portion of the wing experiences a recompression of
tics for the same geometry. The design of a wing the flow prior to the airfoil maximum thickness line;
within this range of p cot A would provide the o p this results in more positive pressures acting on a
portunity to achieve good aerodynamic performance. forward-facing wing surface. On the other hand, the
A review of the lift-curve-slope data of figure 20 outboard aft portion of the wing is characterized by
and the computed drag-due-to-lift results of figure 27 a rearward-sloping surface which combines with the
supports the range of cot A from 0.6 to 0.8 as a fea- high negative pressure coefficients to produce high
sible design region. The data of figure 20 show that drag levels. The other two quadrants of the wing up-
at a value of p cot A of 0.6 the data for the sharp per surface would have pressure fields which combine
leading-edge wings level off and at a value of /3 cot A favorably with the local surface geometry to produce
of 0.8 the same occurs for the blunt leading-edge drag reductions. These observations suggest that im-
data. For flat wings, the lift-curve slope is inversely proved supersonic performance could be achieved if
proportional to the drag-due-to-lift characteristics; the wing designer configures the wing geometry to
therefore, a leveling off of the lift efficiency should take advantage of the natural conical flow structure

17
against it. Future supersonic wing designs should lower surface normal-force coefficient (c&2 c&),
break from tradition and begin to explore new airfoil the data presented in figures 59 and 60 can be
shapes and distributions of airfoils, such as increas- used in an iterative sense to define a feasible high-
ing leading-edge bluntness, airfoil maximum thick- lift separated-flow design space. To assist in the
ness, and airfoil maximum thickness position with iteration process, plots of MN and Q N against Q are
increasing spanwise position to take full advantage presented in figure 61. Presented in figure 62 is a
of the conical nature of the flow. matrix of points, identified by their associated Mach
To define a high-lift (0.2 5 C, 5 0.4) wing de- number and leading-edge sweep values, that satisfy
sign philosophy, an understanding of the flow lim- the distributive normal-force coefficient relationship
itations which exist is required. At these high-lift (C& 2 C&) for a design condition of CN x 0.40.
conditions, the flow will undoubtedly separate at the Each of these points in the matrix was determined
leading edge for an uncambered subsonic leading- by iterating through figures 59 and 60 until the
edge wing independent of airfoil shape. The lee- distributive requirement was satisfied. A sample
side flow condition which would exist for delta wings design iteration for CN = 0.4 and A = 75' is initiated
at angle of attack is shown in figure 17. The data by determining the Q N at which C& = 0.2 and
of figure 17 clearly show that within the feasible then through trial and error the maximum Mach
range of values of p cot A identified for low lift number is determined at which C& = 0.2. Once
(0.4 5 p cot A 5 0.8), MN < 1.00, the lee-side flow O N and M are known, the graphs in figure 61 can
is characterized by a leading-edge separation. By be used to compute M N for plotting in figure 62.
limiting the high-lift design to separated flows and The matrix of feasible solutions presented in figure 62
by imposing the minimum supersonic Mach number can be thought of in terms of lines of constant Mach
identified in figure 28 and an arbitrary selection of number (solid lines) in which both Q and A vary and
A = 75' as the maximum sweep angle, the M N and lines of constant leading-edge sweep (dashed lines)
Q N space can be reduced significantly. (See fig. 58.) in which only M varies. The boundaries of a given
The primary premise on which the high-lift wing de- design space are defined on the right by the maximum
sign concept is to be based is the assumption that wing leading-edge sweep under consideration, on the
the lift force on the wing upper surface must be equal bottom by the minimum Mach number, and on the
to or greater than the lift force on the lower surface. left by the minimum allowable wing leading-edge
The supporting argument for these criteria is that the sweep under consideration. The upper boundary is
negative pressure coefficients acting on the upper sur- defined by iterating through the design process to
face at the leading edge of the wing are the primary determine the maximum Mach number for which a
mechanism for creating aerodynamic thrust; thus, if given wing geometry (leading-edge sweep) satisfies
the upper surface of the wing does not produce at the distributive requirement. Shown in figure 62 are
least 50 percent of the lift, then significant drag re- the maximum Mach numbers for several leading-edge
ductions may not be realized. In order to satisfy sweep conditions.
these conditions, delta wing upper surface and lower
Presented in figure 63 is the design space for de-
surface normal-force coefficient design graphs have
sign values of CN of 0.40 and 0.20. The data indicate
been developed from the data of figures 34 and 36
that a reduction in the design value of CN expands
(figs. 59 and 60). The upper surface normal-force
the range of feasible solutions and shifts it to a lower
coefficient characteristics are strongly dependent on
value of Q N and MN . A decrease in design CN allows
Mach number; for example, an increase in Mach
for the extension to higher Mach numbers. The a p
number from 1.50 to 2.00 reduces the upper surface
parent reduction in the size of the design space with
normal-force coefficient by 50 percent. In a similar
reduced design CN is strictly graphical in nature due
manner the lower surface normal-force coefficient is
to the nonlinearly decreasing relationship between a
a strong function of leading-edge sweep. Additional
and Q N . The design-space concept for moderate-
data show that both the upper surface and lower sur-
to-high-lift conditions (0.2 I C, 5 0.4) was evalu-
face normal-force coefficients of delta wings are not
ated with data from previous supersonic wing designs
significantly influenced by thickness or camber; thus,
(ref. 39). The study showed that designs conducted
the application of these graphs to the development of
for conditions which lie within the appropriate design
a high-lift wing design space for delta wing geome-
space had lower drag due to lift compared with those
tries should be acceptable.
outside the design space.
By imposing a design CN value and requiring that
the design CN can be distributed between the upper To further evaluate the design-space concept and
and lower wing surfaces such that the upper surface to relate the high-lift requirements to the low-lift
normal-force coefficient always equals or exceeds the requirements, lines of constant values of p cot A

18
of 0.4, 0.6, and 0.8 are presented in each plot. The the camber are integrated together to take full ad-
value of O N for each point in figure 63 corresponds to vantage of the conical flow characteristics over the
the value of a~ at which C& equals 50 percent of the wing at zero-lift to high-lift conditions, then aerody-
total design CN value for each wing sweep. The value namic performance levels which exceed the practical
of M N was then determined by iterating through fig- goal limits might be realized.
ure 61(a) for the particular value of @ cot A. The
graphs show that decreasing both @ cot A and the Concluding Remarks
design CN expands the design matrix. For the de-
Through the empirical correlation of experimen-
sign CN of 0.4, decreasing @ cot A from 0.8 to 0.6
tal data and theoretical analysis, a set of graphs have
and 0.4 increases the maximum wing leading-edge
been developed which quantify the inviscid aerody-
sweep for designs from 55' to 65' and to a value
namics of delta wings at supersonic speeds.
greater than 75', respectively. Similarly decreas-
ing design CN to 0.2 increases the maximum wing The zero-lift wave-drag characteristics of delta
leading-edge sweep from 55' to 60' for a design at wings with diamond, circular-arc, and NACA modi-
a value of @ cot A of 0.8. These results have been fied four-digit-series airfoils were determined through
summarized in figure 64 in which the limiting sweep the application of a nonlinear computational tech-
angle for a delta wing has been plotted against the nique. The nonlinear analysis varied substantially
leading-edge sweep parameter @ cot A for design val- from the exact linear-theory predictions for all com-
ues of CN of 0.2 and 0.4. This graph clearly shows binations of geometry and flow parameters under
the dominance that the @ cot A condition has on the study. The nonlinear analysis showed that for slender
ability to produce efficient lifting conditions required wings with highly subsonic leading edges the zero-
for high-lift wing design. lift wave-drag correlation relationship is maintained;
however, as the wing geometry becomes nonslender,
If the high-lift design trade data of figure 64 are the flow about the wing becomes nonlinear and the
combined with the zero-lift data of figure 13 and the relationships which define the zero-lift wave-drag cor-
low-lift data of figures 20 and 27, it may be con- relation parameter are not maintained.
cluded that a value of p cot A of approximately 0.6
composed of A = 65' and M = 1.63 would pro- The aerodynamic characteristics of delta wings at
vide the optimum performance for lift coefficients lifting conditions have been evaluated for the effects
between 0.0 and 0.4. For a conventional design, of wing leading-edge sweep, leading-edge bluntness,
the wing would be configured with a moderately and wing thickness and camber and then summa-
thick airfoil ( t / c x 0.04) with maximum thickness rized in the form of graphs which may be used to
located at 20 percent of the chord to minimize the assess for the aerodynamics in the preliminary de-
zero-lift drag at @ cot A = 0.6 and to maximize the sign process. Empirical curves have been developed
low-lift drag-due-to-lift characteristics. (See fig. 26.) for the lift-curve slope, nonlinear lift effects, maxi-
The 65' swept delta wing with aspect ratio of 1.86 mum lift, longitudinal stability, and distribution of
I should provide a minimum zero-lift drag penalty and lift between the upper and lower surfaces of a wing.
a significant improvement in drag due to lift com- In addition, the impact of various airfoil parame-
pared with a more slender geometry. At M = 1.63, ters, wing leading-edge sweep, and lift coefficient on
the effect of vacuum limit is minimal, providing a the drag-due-to-lift characteristics has been shown
70-percent increase in upper surface lifting potential theoretically.
over the 75' design (@ cot A = 0.45) and an 8-percent
The various graphs which detail the aerodynam-
decrease from the 55' design (@ cot A = 0.8). If a
ics of delta wings at both zero-lift and lifting con-
simple variable camber system (simple flap), which ditions were then employed to define a preliminary
is conical in nature, is added to the described de-
wing design approach in which both the low-lift and
sign, the aerodynamic performance across the range high-lift design criteria were combined to define a
of lift coefficient could approach that of the practical
feasible design space.
goal established in reference 40. (See fig. 65.) Also
presented in the figure are the flat-wing aerodynamic
characteristics for delta wings with A = 1 and 2 from
figure 27. The projected performance levels proba- NASA Langley Research Center
bly represent the minimum and not the maximum Hampton, Virginia 23665-5225
allowable performance levels. If both the airfoil and December 11, 1987
Appendix
Nonlinear Methodology
The nonlinear computational technique (ref. 2) selected for analysis solves the nonconser-
vative finite-difference analog of the full-potential equation in a spherical coordinate system.
The method marches in the radial direction to obtain three-dimensional cross-flow solutions of
the wing geometry. The ability to compute the flow over a wing with this method is restricted
to conditions in which the flow in the marching direction remains supersonic; thus, the code
would be expected to be more successful for higher free-stream Mach numbers. The solution
technique is graphically depicted in figure A l . A series of three transformations are used in the
code: the geometry in Cartesian coordinates is transformed to spherical coordinates followed
by a stereographic projection (not shown) and a conformal mapping to a circle. A shearing
transformation is then performed to the computational plane. Grid points are internally com-
puted within the code and are positioned between the body surface (inner boundary) and the
bow shock (outer boundary). The method has been shown to be ideally suited for the solution
of three-dimensional wing flow fields (ref. 2).
Solutions were obtained by fitting the bow shock and capturing all internal shocks. Due to
the supersonic nature of the wing trailing edge, the wake was modeled as an extension of the
wing surface. Sensitivity studies were conducted to determine marching step size and cross-
plane grid resolution. These studies indicated a dependence of both parameters to free-stream
flow conditions as well as geometry. Selection of the marching step size was based upon the
analysis of a series of airfoil geometries at various flow conditions. A representative sampling
of this analysis is presented in figure A2 in which the axially integrated drag values for three
airfoil types are presented for step sizes of 0.2, 0.5, and 1.0, which correspond to 50, 20, and 10
marching steps, respectively. Results of the analysis show a convergence of the solution with
decreasing step size. Based upon these results and considerations of computational cost, a step
size of 0.5 was selected for the study. Cross-plane mesh density was selected to be 30 x 30 based
upon a similar parametric study. Presented in figure A3 are representative 30 x 30 cross-plane
computational grids for the four basic airfoil types under investigation: diamond, circular arc,
sharp NACA modified four-digit series, and blunt NACA modified four-digit series. The grid
plots clearly show clustering of the grid points at the wing leading edge in both the radial and
circumferential directions; this is critical to resolving the wing flow field.

NCOREL A = 1.O delta wing


circular-arc airfoil

‘t
(X/Cr = 0.6)

Physical plane

Mapped plane

Computational plane

Figure A l . Nonlinear method solution process used for delta wing study.

20
'D.W
.0 1 Ci rcular-arc

=:[[ ai rfoi I DX Four-digit series Circular arc Diamond


0.2 0.0032 0.0030 0.0025
0.5 0.0029 0.0027 0.0023
cDiw (DX) 1.0 0.0025 0.0024 0.0020

cD,W (DX)
"

.O1
Diamond airfoil
,[[ Four-digit-series

/
0 1.o 0 1.o
C
'r XJCr
Figure A2. Selection of step size for analysis based upon delta wing with A = 1.0 and
airfoils with T = 0.04 and m = 0.5.

r 7
Diamond airfoil
=0.04. m = 0.5 f Circular-arc airfoil

Modified four-digit- Modified four-digit-


series airfoil series airfoil
7 = 0.04, m = 0.5,R = 0

Figure A3. Representative computational grids for A = 1.0 delta wing at M = 1.41 and a = 0'.

21
References 16. Kaattari, George E.: Pressure Distributions on Trian-
gular and Rectangular Wings to High Angles of Attack-
1. Weber, J.; and King, C.: Analysis of the Zero-Lift Wave Mach Numbers 1.45 and 1.97. NACA RM A54D19,1954.
Drag Measured on Delta Wings. R. & M. No. 3818,
17. Hall, Charles F.: Lift, Drag, and Pitching Moment
British Aeronautical Research Council, 1978.
of Low-Aspect-Ratio Wings at Subsonic and Supersonic
2. Siclari, Michael J.: The NCOREL Computer Program for
Speeds. NACA RM A53A30, 1953.
3D Nonlinear Supersonic Potential Flow Computations.
18. Boatright, William B.: Experimental Study and Analysis
NASA CR-3694, 1983.
of Loading and Pressure Distributions on Delta Wings
3. Hill, William A., Jr.: Experimental Lift of Low-Aspect-
Due to Thickness and to Angle of Attack at Supersonic
Ratio Triangular Wings at Large Angles of Attack and
Speeds. NACA RM L56114, 1956.
Supersonic Speeds. NACA RM A57117, 1957.
4. Love, Eugene S.: Investigations at Supersonic Speeds of 19. Briggs, M. M.; Reed, R. E.; and Nielsen, J. N.: Wing-
22 Triangular Wings Representing Two Airfoil Sections Alone Aerodynamic Characteristics to High Angles of
for Each of 11 Apex Angles. NACA Rep. 1238, 1955. Attack at Subsonic and Transonic Speeds. NEAR T R 269
(Supersedes NACA RM L9D07.) (Contract DAAG29-79-C-0020), Nielsen Engineering &
5. Henderson, Arthur, Jr.: Experimental Investigation of Research, Inc., Nov. 1982. (Available from DTIC as AD
the Zero-Lift Wave Drag of Seven Pairs of Delta Wings A125 764.)
With Constant and Varying Thickness Ratios at Mach 20. Michael, William H., Jr.: Flow Studies on Flat-Plate
Numbers of 1.62, 1.99, and 2.41. NACA RM L55D13, Delta Wings at Supersonic Speed. NACA TN 3472, 1955.
1955. 21. Miller, David S.; and Wood, Richard M.: Lee-Side Flow
6. Stallings, Robert L., Jr.; and Lamb, Milton: Wing-Alone Over Delta Wings at Supersonic Speeds. NASA TP-2430,
Aerodynamic Characteristics for High Angles of Attack at 1985.
Supersonic Speeds. NASA TP-1889, 1981. 22. Boyd, John W.; and Phelps, E. Ray: A Comparison
7. Smith, Fred M.: Experimental and Theoretical Aero- of the Experimental and Theoretical Loading Over Trian-
dynamic Characteristics of Two Low-Aspect-Ratio Delta gular Wings at Supersonic Speeds. NACA RM A50J17,
Wings at Angles of Attack to Sa" at a Mach Number 1951.
of 4.07. NACA RM L57E02, 1957. 23. Britton, J. W.: Pressure Measurements at Supersonic
8. Hatch, John E., Jr.; and Hargrave, L. Keith: Eflects Speeds on Three Uncumbered Conical Wings of Unit As-
of Reynolds Number on the Aerodynamic Characteristics pect Ratio. C.P. No. 641, British Aeronautical Research
of a Delta Wing ut Mach Number of 2.41. NACA Council, 1963.
RM L51H06, 1951. 24. Michael, William H., Jr.: Flow Studies on Drooped-
9. Gallagher, James J.; and Mueller, James N.: A n In- Leading-Edge Delta Wings at Supersonic Speed. NACA
vestigation of the Maximum Lift of Wings at Supersonic TN 3614, 1956.
Speeds. NACA Rep. 1227, 1955. (Supersedes NACA 25. Igglesden, M. S.: Wind Tunnel Measurements of the
RM L7J10.) Lift-Dependent Drag of Thin Conically Cambered Slender
10. Squire, L. C.; Jones, J . G.; and Stanbrook, A.: A n Delta Wings at Mach Numbers 1.4 and 1.8. Tech. Note
Experimental Investigation of the Characteristics of Some No. Aero 2677, British Royal Aircraft Establishment,
Plane and Cambered 650 Delta Wings at Mach Numbers Apr. 1960.
From 0.7 to 2.0. R. & M. No. 3305, British Aeronautical 26. Abbott, Ira H.; and Von Doenhoff, Albert E.: Theory of
Research Council, 1963. Wing Sections. Dover Publ., Inc., c.1959.
11. Ellis, Macon C., Jr.; and Hasel, Lowell E.: Prelimi-
27. Bishop, R. A.; and Cane, E. G.: Charts ofthe Theoretical
nary Investigation at Supersonic Speeds of Triangular and
Wave Drag of Wings at Zero-Lift. C.P. No. 313, British
Sweptback Wings. NACA TN 1955, 1949.
Aeronautical Research Council, 1957.
12. Vincenti, Walter G.; Nielsen, Jack N.; and Matteson,
28. Puckett, Allen E.: Supersonic Wave Drag of Thin Air-
Frederick H.: Investigation of Wing Characteristics at a
foils. J. Aeronaut. Sci., vol. 13, no. 9, Sept. 1946,
Mach Number of 1.59. I-Triangular Wings of Aspect
pp. 475-484.
Ratio 2. NACA RM A7110, 1947.
13. Ulmann, Edward F.; and Dunning, Robert W.: Aero- 29. Middleton, W. D.; Lundry, J. L.; and Coleman, R. G.:
dynamic Characteristics of Two Delta Wings at Mach A System for Aerodynamic Design and Analysis of Super-
Number 4.04 and Correlations of Lift and Minimum-Drag sonic Aircraft. Part 2-User's Manual. NASA CR-3352,
Data for Delta Wings at Mach Numbers From 1.62 to 6.9. 1980.
NACA RM L52K19, 1952. 30. Wood, Richard M.; and Miller, David S.: Ezperimen-
14. Hatch, John E., Jr.; and Gallagher, James J.: Aero- tal Investigation of Leading-Edge Thrust at Supersonic
dynamic Characteristics of a 68.4' Delta Wing at Mach Speeds. NASA TP-2204, 1983.
Numbers of 1.6 and 1.9 Over a Wide Reynolds Number 31. Miller, David S.; and Wood, Richard M.: An Inves-
Range. NACA RM L53108, 1953. tigation of Wing Leading-Edge Vortices at Supersonic
15. Menees, Gene P.: Lift, Drag, and Pitching Moment of Speeds. AIAA-83-1816, July 1983.
an Aspect-Ratio-2 Triangular Wing With Leading-Edge 32. Mayer, John P.: A Limit Pressure Coeficient and an
Flaps Designed To Simulate Conical Camber. NASA Estimation of Limit Forces on Airfoils at Supersonic
MEMO 10-5-58A, 1958. Speeds. NACA RM L8F23, 1948.

22
33. Stanbrook, A.; and Squire, L. C.: Possible Types of Flow 37. Mack, Robert J.: Wind-Funnel Investigation of Leading-
at Swept Leading Edges. Aeronaut. Q., vol. XV, pt. 1, Edge Thrust on Arrow Wings i n Supersonic Flow. NASA
Feb. 1964, pp. 72-82. TP-2167, 1983.
34. Brown, C. E.; and Michael, W. H., Jr.: Effect of 38. Covell, Peter F.; Miller, David S.; and Wood,
Richard M.: An Evaluation of Leading-Edge Flap Per-
Leading-Edge Separation on the Lift of a Delta Wing. J.
Aeronaut. Sci., vol. 21, no. 10, Oct. 1954, pp. 690-694, formance on Delta and Double-Delta Wings at Super-
706. sonic Speeds. AIAA-86-0315, Jan. 1986.
39. Miller, David S.; and Wood, Richard M.: Aerodynamic
35. Squire, L. C.: Camber Effects on the Non-Linear Lift of Design Considerations for Efficient High-Lift Supersonic
Slender Wings With Sharp Leading Edges. C.P. No. 924, Wings. AIAA-85-4076, Oct. 1985.
British Aeronautical Research Council, 1967.
40. Wood, Richard M.; Miller, David S.; Raney, David L.;
36. Squire, L. C.: Experimental Work on the Aerodynamics and Roesch, Michael T.: A Low-Lift Wing Camber De-
of Integrated Slender Wings for Supersonic Flight. Prog. sign Approach for Fighter Aircraft. NASA TP-2465,
Aerosp. Sci., vol. 20, no. 1, 1981, pp. 1-96. 1985.

23
Table I. Experimental Data Set

Planform Flow conditions


teference A 5Pe tI C LE 7orce Pressure Flow visualizatio1
3 3 wings Modified 4% at Sharp J
76' to 84.6' biconvex 0.59~ 4.48 x lo6
4 11 wings Diamond 8% at Sharpand J
45' to 80' 0.18~ blunt 1.39 x lo6

:.; 1 1
4 wings Slab Sharp and J
65' to 75' blunt I6 1 1 . 3 9 ~lo6
5 4 wings Diamond 6% at Sharp 1.62 to 1-2 to 12.36 x lo6 to J
48' to 73' 0.20c to 11.32 x 10;

1.60 to -5 to 1.41 x 10 to
2.82 x lo6
I I
4.07 1-4 to 15.3 x IO6

150 I
68.6'

45' and 64'

10 2 wings
65'
11 8 wings
36.4' to
70.6'
12 2 wings
63.43'

15 1 wing NACA-0003 6% Blunt 1.30 to -6 to 3.68 x 10'


63.26' 2.22 18
16 2 wings Biconvex 5% at Sharp 1.45 and 0 to 0.68 x lo6 to
45' and 0.50~ 1.97 50 3.27 x lo6
63.4'
17 3 wings NACA-0003 3 to 8% Sharp a d 1.2 to -4 to 1.4 x lo6 to
45' to md Biconvex blunt 1.7 20 3.0 x lo6
63.26'

24
3
F
Planform Airfoil Flow conditions Data t v w
Reference A Type 0 , deg NRe/E Force Pressure Flow visualization
3 wings 65A003 Blunt 1.62 to 0 to
53' to 20
66.6'

53O

3 wings
Flat

Slab 0 to 1 x 106 to
I 1 I d
63.43' to
82.87' I /P20
55 2 x 106

20 6 wings
58.25' to
85'
4 wings
52.5' to
Flat

Flat
Sharp 1.90

Sharp 1.50 to
2.80
0 to
20

0 to
20
5.21 x 10' to
9.38 x lo6

0.86 x lo6 to
2.46 x lo6
d

d d
75'
2 wings NACA 000663 6% Blunt 1.20 0 to 1.8 x 10' d d
45O 1.30 to 20
45' Biconvex Sharp 1.70
3 wings Cross sections: Conical Sharp 1.30 to 0 to 3.63 x lo6 to ,/ d

I 12.Bo
76' 30' Diamond 16 2.26 x lo6
30' Biconvex
60' Biconvex
3 wings Flat 0 to
58.25O to Cambered 20
75'

I 25
8 wings
71.56'
Slab
Cambered
-8 to
12
0.66 x 106 d

25
5

3
A
2

0 10 20 30 40 50 60 70 80 90
A , deg
Figure 1. Relationship between A and A for delta wings.

65"

55"
M
45"

0 .2 .4 .6 .8 1.0 1.2 1.4 1.6 1.8 2.0


p cot A
Figure 2. Relationship between M and p cot A for delta wings.

26
e C I

m = rlc
z =uc
S = cbl2 b

1
A=b2/S

Figure 3. Identification of geometric parameters for delta wings with diamond airfoil.

27
Circular-arc airfoil - thickness

I
0
X/C
1.o

.05

0 1 .o
X/C

m Diamond airfoil - thickness location ( z = 0.04)

* 2 0 *
0 1.o
X/C
Figure 4. Representative profiles of circular-arc and diamond airfoils.

20
Thickness (m = 0.4, R = 6)
.OE e

0 1.o
xlc

Maximum thickness location ( z = 0.04, R = 6)

0 1 .o
X/C

Leading-edge radius ( z = 0.04, m = 0.4)

0 1.o
X/C

Figure 5. Representative profiles of NACA four-digit-series airfoil.

29
0 v)
7

30
9
7

W
1
5
x

1 9
r

w
1
5
x

9
.-

W
-I
,x
x

9
7

W
-1
P
x

.o :o
0
Lo 0 Lo Lo 0 Lo
Q Q
0 0

II E
E

31
\ I

i
\
\
\
\
\
\
1

o m
5 m b
0 0 0
a.

'I
I \
' I
I , '\
I I I
0

0
7

0
a9

32
h
0
II
LT:

0
II

I
u!
h
X
W

‘Y
0
9 h
X
W

33
m !05
F W

34
i
m
/

c.

/
7

9
7

II
a

35 I
6-
000
000
000

U 99
o*a

\\
G
w-. *
- *b--- \

v
2 cum.
00
3 99
c 00
.-
0 d
.-
c 0
v)
0
QO
$ II E cuL.r!
00
i
a,U t
t I
Y
.-0 I
E I
I
I
I
I
I
0

h
X
Y h
X
Y

v)

36
t

i 0
z
G
V
v! 4
E
a

37
/ m

o m
5 m b
0 00
c? a
I I d
I I 9
00

I0
9
I
II II
ru

0
7

OLD
-00
II 11 It
a E u
0
0 09
T.

38
t

c m o m a3
cvu3r'.
"
0
000

cue a3
P 00
00
I
I
I U
I
I
I
I
0

E 2: - m
\
I
I

1
1
/
I
I
I
I
I

3
U

39
- 1.o
1.0

-.5 - / \

z
1.6,- R=o

cP O --
-1.4 .5 L
2
Z

-1.4
0
YlYLE
c p .5
1.9
O
b 0
%€
1.o 0
Y%€
1.o 0
Y%€
1.o 0
YIYLE
1.o

Figure 16. Effect of bluntness on spanwise surface pressure distributions predicted by nonlinear theory for
delta wings with A = 1.0, 7 = 0.04, and m = 0.5 at M = 3.16.

40
2.00 I-

Classical vortex

1 0 10 20 30 40 50
I
aN
Figure 17. Lee-side flow classification chart for delta wings (ref. 29).

41
A = 75". M = 2.00 A=75O, a = 8"

.4
0 5 10 15 20 1.50 2.00 2.50 3.00
a,deg M

1.0 - Vortex
.9 - action line

.8 -
rl, .7
.6
-8,
-
77331qv
.5 -
Wings with
.4
I I I zero thickness

Figure 18. Location of vortex action line for zero-thick delta wings (ref. 31).
2
0
a
5 9
m
2
a
.-c

QbP 0.

IA 9
cu

\
4a

9
7

43
.08 NCOREL Airfoil
----- Sharp
Linear
theory
Experiment
blunt airfoil
Ex pe ri me nt
sharp airfoil
PCL '
a .04
deg-'

0 .2 .4 .6 .8 1.o
p cot A
Figure 20. Nonlinear-theory-predicted lift characteristics of thin delta wings.

44
-
ACL= C - 20 c
I
L a 0"
.2 Airfoil
Linear Blunt Sharp Reference
3 14
I lift I 6
I I h
32
A
v
I I P
I I
I
U 14
I
D I I 9 15
I
I
I
I
I
II
I
I
I
' 0
A
V
16
17
18
0 19
ACL 0
at a = 20" i
I I
I I
I
I I v-7
I
I
I
I
.u.

I I

-.2
0 1.o 2.0
4p CO$ A
Figure 21. Nonlinear lift effects of thin delta wings.

45
1.;

w
.&
U
D

‘L, max
Sharp
airfoil Reference
6
.4 0 7
D 32
[1 9
0 16

0 5 10
WA
Figure 22. Maximum lift coefficient of thin delta wings.

I
46
O a
4 Q

4 QQ
a

v
\ I
I
.- \

a=

0
1%
I Q
4 a
Q

47
a
2

b
v
\ I
\ I

8
0
7

E
0
II 0 0
E
0 0
a 0
0
0
0 0
0 0
0
0
I
e 0
0

48
.68 .44

.42

Ycp W 2 )
Xcp/Cr -64 .40
at a = 5" at a = 5"

.38

.60 .36
0 1.o
M/A

(a) Q = 5'.

.68 .36

.34

Xcp/Cr Ycp /(b/2)


.64 .32
at a = 50" at a = 50"

30

.60 28
0 5 1.o
M/A

(b) Q = 50'.
Figure 25. Wing panel center-of-pressure location for thin delta wings (ref. 6).

49
P '

50
.7 Linear-theory Sonic
zero thrust LE
NCOREL CL A
0.1 1
0.2 1
0.1 2
0.2 2

.5

Linear
theory
.3 -full thrust

.1
0 .2 .4 .6 .8 1.o

P cot A
Figure 27. Nonlinear- and linear-theory-predicted drag-due-to-lift characteristics of delta wings.

51
0 .2 .4 .6 .a 1.o
rl

(a) M = 0.60 (ref. 19).

Figure 28. Spanwise pressure distributions of delta wing with A = 1.0 at a = 15'.

52
-1 .o

-.8
dcr
0 0.500
0 0.625
0 0.750
-.6 A 0.875

-.4

0
1 A A

<
t -21

.4

.6 1 I I I I
0 ,2 .4 .6 .8 1.o
rl

(b) M = 0.90 (ref. 19).

Figure 28. Continued.

53
-1 .o

-.8

-.6

0 0.500 A
0 0.625 OA
W o0+ l
-.4
0
A
0.750
0.875 n
0
w

cp -.2
e

.4

.6
0 .2 . .4 .6 .8 1 .o
rl

( c ) A4 = 1.20 (ref. 19).


Figure 28. Continued.

54
I
-1 .o

-.8

-.6
VCr
0 0.500
0 0.625
0 0.750
-.4

0
-.2

CP 0

.2
4

.4

.6
0 .2 .4 .6 .8 1.o
I
rl
(d) M = 1.60 (ref. 6).

Figure 28. Continued.

55
-1 .o

-.8

-.6
dcr
0 0.500
0.625
0 0.750
-.4 A 0.875

cp -.2

c
.L

.L

.€
0 .2 .4 .6 .8 1.o
rl

(e) M = 2.16 (ref. 6).

Figure 28. Continued.

56
-1 .o

-.8

-.6
VCr
0 0.500
0 0.625
0 0.750
-.4 A 0.875

cp -.2

4
0

.a

.4

.6
0 .2 .4 .6 .8 1.o
rl
(f) M = 2.86 (ref. 6).

Figure 28. Continued.

57
-1.o

-.8

-.6
dcr
0 0.500
0 0.625
0 0.750
-.4 A 0.875

-.2

.4

.6
0 ,2 .4 .6 .8 1.o
rl

(g) M = 3.50 (ref. 6).


Figure 28. Concluded.

58
1.(

M R ference
-.t - 0 0.60 19
0 0.90 19
0 1.20 19 n
A 1.60
h 2.16
- b 2.86
-.E a 3.50

-.4

-.2

cP O

.2

.4

0 .2 .4 .6 .8 1 .o
rl
Figure 29. Effect of Mach number on spanwise pressure distributions of delta wing with A = 1.0 at a = 15'
and x / c T = 0.625.

59
Reference a ,deg M

-1.2 - 19
10
-a
20
__d 0.60

-.6

cP
z
b
0
k
f
.6
0 .2 .4 .6 .a 1.O
rl

Figure 30. Effect of angle of attack on spanwise surface pressure distributions of delta wing with A = 1.0
for z / c r = 0.625.
I I Upper surface
v7A Lower surface

M = 0.60 M = 0.90 M = 1.20 M = 1.60


100

80

60
Percent
of lift
40

20

0
10 20 10 20 10 20 10 20

i
Figure 31. Distribution of lift for delta wing with A = 10.

61
a degt

10A 20
0 0’ 0.5
0 1.0 d
0 2.0
Upper surface
a
------ Lower surface

cP

.6 I I I I I
0 .2 .4 .6 .a 1.o
rl
(a) M = 0.60 (ref. 19).

Figure 32. Effect of A on spanwise surface pressure distributions of delta wings at z / c r = 0.750.

62
a deg
I

10 A 20
-1.2 0 d 0.5
0 1.0 d
0 2.0 0
Upper surface
------ Lower surface

-.6

U-

.6
0 .2 .4 .6 .a 1 .o
rl

(b) M = 1.60 (ref. 6).

Figure 32. Concluded.

63
2
c
0

0 0 0 0 0 0
0 a3 (D d (N

64
M
.6
1.60 2.16 2.86 3.50 A, deg A
0 (> 82.87 0.5
0 d rn 75.96 1.0
o d 63.43 2.0

0
0
.4
0
00
no
0
0
0 d cr a
.2
0 d*
OO
a
a

0 10 20 30 40 50
a p cot A
N
(a) Reference 6.

Figure 34. Wing upper surface normal-force coefficient for thin delta wings.

66

!
M
.6
1.50 1.70 2.00 2.40 2.80 4 deg A
0 0 0 75 1.07
m
8
n
67.5
60
52.5
1.66
2.31
3.07

OA
.4
A
0
n
d

.2
Q Q
Od
n 0 0
0 0
0

0 10 20 30 40 50
aNpcot A

(b) Reference 21.

Figure 34. Continued.

68
.6 M = 1.90
A, deg A
0 85 0.35
82.5 0.53
.5 0 80 0.71

2n 75
67.5
58.25
1.07
1.66
2.48

.4

A
.3
A
CUN
i
41!
d,
A
.2

A
00
.1

0 10 20 30 40 50
aNp cot A

(c) Reference 20.

Figure 34. Concluded.

07
-.6 M
1.60 2.16 2.86 3.50 A , deg M = 1.60
0d a
0 82-87 ‘//// cp,v
0d CB W 75.96
0 0 63.43 0 90% cp,v
0 0
e
-.4 0

U
Cp, min
0
cl
0
0

cr
-
a90%
M = 2.16
cp,v
c p,v

-.2
00

ca
dcb
- M = 2.86
cp,v

06
(>
0 .

(a) Reference 6.

Figure 35. Minimum upper surface pressure coefficient for thin delta wings.
M
1.50 1.70 2.00 2.40 2.80 A , deg
0 d c f @ . 7 5
67.5
o d
O G $ o 60
nLY 52.5

-.6

0
0
0 - M = 1.70
CPY

U
Cp, min
-.4 n

% d 90% c p , v
- M = 2.00
cp,v

0
ON “Q
6
M = 2.40
‘//// cp,v
90% cp,v
-.2
= 2.80
88% Cp,v C P,V
0 0
0

I I I 1 I
0 10 20 30 40 50
a~ p Cot A

(b) Reference 21.

Figure 35. Continued.

69
M = 1.90
-.6
A , deg
0 a5
0 82.5
0 80
-.5 A 75
A 67.5

-
L3 58.25

M = 1.90
-.4 CPJ

U
A
Cp, min -3 A
A

-.2

-.1

(c) Reference 20.

Figure 35. Concluded.

70
00
00
00
0 0
0
n o

00
00 4
*p8

00 h 4
7

U
7
0

U 4

000

71
-.8
Airfoil
Sharp
----- Blunt

-.6

= 1.60

U
C p, min - - 4
= 1.90

= 2.40

-.2

0 10 20 30 40 50
a N p Cot A
(a) Summary plot.

Figure 37. Effect of wing leading-edge bluntness on C;,,min for delta wings.

72
M
1.62 1.94 2.41 A , deg Airfoil
-.6 0 53.0 Sharp
0 53.0 Blunt
0 c> 60.0 Blunt
A A A 66.6 Blunt

-.4
0 0
A
I> I> A
U A A
Cp, min 0
A
A
I> A A A
-.2

0 10 20 30 40 50
CY. N p Cot

(b) Reference 18.

Figure 37. Continued.

73
0
-.6 0
0

-.4 0

0 0
U 0
Cp, min 0 0

0
-.2 0
A = 63.5"
M
0 1.45
n 1.97

0 20 30 40 50
CX N p Cot A

(c) Reference 16.

Figure 37. Concluded.

74
Reference M Re
0 8 2.41 lo6 to 18.3x lo6
0 14 1.62 7.2 x l o 6
0 14 1.62 2.4 x l o 6
0 14 1.93 7.2x l o 6
+ 14 1.93 2.4 x l o 6
-.E
+ 14 1.93 12.6x lo6
cr 14 1.93 18.4x lo6

0
0
0 0
-.4
n

U
Cp, min

-.2 0 0
0 0 0 0

0 10 20 30 40 50
CXN p Cot h
~ Ci,,minfor delta wings with A = 68.6'.
Figure 38. Effect of N R on

75
.4
0 5 10 15 20 1.50 2.00 2.50 3.00
a,deg M

M = 2.00, a = 8" Wings with


zero thickness
-4- Cross section

rl,
.8
.7
8

'
0 y o
/

.6
0 y>-,
f

.4
50 60 70 80
4 deg
Figure 39. Location of vortex action line for thick delta wings (ref. 23).

76
co
0 *
0 0

77
c
.L

at a = 16'

-.2
0 1.o 2.0
4 p cot2 A
Figure 41. Nonlinear lift effects of thick delta wings (ref. 23).

78
Cross section
-.3 -

-.2

-.1 -
------ Lower surface

cP 0

.1

.2

.3
0 .2 .4 .6 .8 1.o
rl

Figure 42. Effect of thickness on spanwise pressure distributions of thick delta wings at M = 1.30
and a = 8.0' (ref. 23).

79
-.3
M
0 1.30
1.60
$A 2.00
2.40
-.2
A 2.80

Upper
Lower

-.1

.1

1
4

.2

.3
0 .2 .4 .6 .8 1.o
rl

Figure 43. Effect of Mach number on spanwise surface pressure distributions for thick delta wings
at a = 8.0' (ref. 23).

80
Upper surface
----- Lower surface

4
-.,2

cP

-.1

0
P
I

.1

.2

.3

.4.
0 .2 .4 .6 .8 1.o
T

Figure 44. Effect of angle of attack on spanwise surface pressure distributions of thick delta wings
,
at M = 1.30 (ref. 23).

81
A = 1.0
r l Upper surface

0 3 M = 1.30 M = 2.00
Lower surface

M = 2.80

100

80

60
Percent
of lift
40

20

n
8 16 8 16 8 16

a,deg

Figure 45. Distribution of lift for thick delta wings (ref. 23).

82
M
2.00 2.80
d Q
-.6 d I

0 9
Thin wing
-.5 M = 1.30

-.4

CUN -.3

-.2

-.1

0 10 20 30 40 50
a,,,p cot A

Figure 46. Comparison of upper surface normal-force characteristics on thin and thick delta wings (ref. 23).

a3
Thin wing
M
-.6 -
1.30 2.00 2.80
30"
0
= .01
d
.07 .06
a
-.5 - 30" 0 d 0
= .07 .05 .04

I" 60"
0
= .15
a +
.10 .10
a =12"
-.4

U
M = 2.00
Cp, min - a 3

-.2

-.l

0 10 20 30 40 50
aN p Cot A

Figure 47. Comparison of upper surface minimum pressure coefficient on thin and thick delta wings (ref. 23).
0
a3

0
b

0
to

0
v)

0
d

3
3

3
0 N
"so
cu 0
=cue o b 0
-
3

c?o
0
7
0 0

I I I I I I

85
A=75", M = 2.01 A=75O, a = 8"
q;--q
\ \
\ \

0 5 10 15 20 1.50 2.00 2.50 3.00


a M

1\
M = 2.00, a = 8"
Zero thickness
1 .o d--4--d
\
M = 1.90
.9
Ay/(b/2) AY
.8 A , deg 0.10 0.20
o d
L
rl, .7 "0 75
67.5 El d
58.25 0
.5
.6
.4 f-
50 60 70 80
l y
A

Figure 49. Location of vortex action line for cambered delta wings (ref. 24).

86
.08

Linear theorv

/
Thin wings

PCLa .04

Reference

/ 0
0
24
10
/ 0 25
f
0 1.o 2.0
PcotA

Figure 50. Lifting characteristics of cambered delta wings.

87
U
$4
7
0

\
\
\
\
\
\ 0
I
I .E
' C
I
I
/
'trr /

88
-.4
Ay/(b/2)
0 0.10
0 0.20
Upper surface
-.3 - - - - - Lower surface

-.2

-.1

.1

.2

.3
0 .2 .4 .6 .8 1.o
rl

Figure 52. Effect of wing camber on spanwise surface pressure distributions for delta wings at M = 1.40
and CY = 8' (ref. 24).
-.4 - A Y = 0.20
-
b/2

A , deg
0 75.00 f-
-.3 - 0 67.50
0 58.25
Hinge
Upper su rface line
------ Lower surface
-.2

-.1

cP
0

.1

.2

.3
0 .2 ,4
,Tl
.6 v 1.o

Figure 53. Effect of A on spanwise surface pressure distributions for cambered delta wings at M = 1.90
and a = 8' (ref. 24).

90
1AY

Upper surface
= 0.2
Lower surface
-.4 J L .
a,deg
0 0
0 8
-.3 A 16

-.2 4

-.1

cP

.1 6

.2
A

.3
0 .2 .4 .6 .8 1.o
rl

Figure 54. Effect of a on spanwise surface pressure distributions for cambered delta wings at M = 1.90 (ref. 24).

91
Ay/(b/2)
0.1 0 0.20 A , deg
0 d 75
d
B 0
67.5
58.25

Thin wing
M = 1.90

P
.L

0 10 20 30 40 50
uNp cot A

Figure 55. Upper surface normal-force coefficient for cambered delta wings (ref. 24).

92
-.6

A y/( b/2)
0.10 0.20 Ax, deg
o d 75.00
-.5
I7 d 67.50 f I ,
0 6 58.25

-.4 Thin wing


M = 1.90

-.2

-.1

I I I I I I
0 10 20 30 40 50
a p cot A
N

Figure 56. Upper surface minimum pressure coefficient for cambered delta wings (ref. 24).

93
94
2.0

:ached flow

lparate!d flow

MN 1.o

0 10 20 30 40 50
aN

Figure 58. Influence of minimum M , maximum wing sweep, and separated flow restriction on design space.

95
.6

.4

.2

0 10 40 50

Figure 59. Delta wing upper surface normal-force coefficient design curves.
0
co

0
(D

0
e

0
cu

97
2 2 112
MN = M cos A (1 +sin a + tan A )

M
10.00 5.00 4.00 3.00 2.00 1.50 1.20
50

40

30
a
20

10

0 .2 .4 .6 .8 1.o
2 1I2
cos A ( l +sin a tan2A)

A , deg
0 45 55 65 75
50

40

30
a
20

10

0 20 40 60 80 100
aN
Figure 61. Variation of M N and CXNwith a.

98
0

99
0
In

0
T
J.

c
5 %0 '4 b. 0
m
0 0 0

0 z
II 8
z
0 0
N

0
7

- 2
0 0
All
0
3Z In
0
c
c
0
u oc! '4 b.
0 0 0
Q
\ 0
e
\
\
\ 0
m
9
0 \
II \ z
z 8
0 \
I 0
N
I
I
0
7

100
CN
0.4 0.2
80

70
Limiting
sweep Feasible
angle design
60
space

50

1 .2 .4 .6 .8 1.o
p cot A
Figure 64. Identification of limiting leading-edge sweep angle.

Practical Goal (ref.40)


- A = 1 flat wing (fig627)
-- A = 2 flat wing (fig.27)
----- Projected-Variable Camber

Percent of
thrust

'\
\
\ '. \
\
\
'.
I I I I
0 .1 .2 n
.3 .4 .5

Figure 65. Identification of practical aerodynamic performance goals.

101
1. Report No. 2. Government Accession No. 3. Recipient’s Catalog No.
NASA TP-2771
1. Title and Subtitle 5. Report Date
Supersonic Aerodynamics of Delta Wings March 1988
6. Performing Organization Code

7. Author(s)
8. Performing Organization Report No.
Richard M. Wood
L- 16212
10. Work Unit No.
3. Performing Organization Name and Address
NASA Langley Research Center 505-61-71-01
Hampton, VA 23665-5225 11. Contract or Grant No.

13. Type of Report and Period Covered


12. Sponsoring Agency Name and Address
National Aeronautics and Space Administration Technical Paper
Washington, DC 20546-0001 14. Sponsoring Agency Code

15. Supplementary Notes

16. Abstract
Through the empirical correlation of experimental data and theoretical analysis, a set of graphs have
been developed which summarize the inviscid aerodynamics of delta wings at supersonic speeds. The
various graphs which detail the aerodynamic performance of delta wings at both zero-lift and lifting
conditions were then employed to define a preliminary wing design approach in which both the
low-lift and high-lift design criteria were combined to define a feasible design space.

17. Key Words (Suggested by Authors(s)) 18. Distribution Statement


Supersonic aerodynamics Unclassified-Unlimited
Delta wings
Experiment
Nonlinear methodology
Wave drag
Subiect Categorv 02
19. Security Classif.(of this report) 20. Security Classif.(of this page) 21. No. of Pages 22. Price
Unclassified Unclassified 104 A06

You might also like