Sie sind auf Seite 1von 24

International Space Planes and Hypersonic Systems and Technologies Conferences 10.2514/6.

2020-2423
March 10-12, 2020, Montreal, Quebec, Canada
23rd AIAA International Space Planes and Hypersonic Systems and Technologies Conference

Combined Capsule/Waverider Configurations


For Boost Glide Missions
Patrick E. Rodi1
Adelante Sciences Corporation, Houston, TX 77586

The boost-glide mission, as the name implies, contains two very distinct portions of the
trajectory. In most boost-glide scenarios, the first segment is described as a boost through
the atmosphere to a suborbital trajectory followed by a re-entry deep into the atmosphere
where a pull-up maneuver occurs. The second segment is a glide segment where the vehicle
slowly bleeds off energy while moving downrange and/or crossrange. The configuration
design problems are quite distinct in these two segments as well. In the first, the high
Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on April 4, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2423

heating rates of the pull-up maneuver requires a design that can tolerate the high dynamic
pressure and convective heating rates experienced. In the second, range is directly related to
the aerodynamic performance by the lift-to-drag ratio. Historical boost-glide vehicles are a
compromise between the two flight segments.
Waverider vehicles are those that ride their self-generated shock wave system. The shock
wave is attached along the geometry’s leading edge effectively isolating the upper (leeward)
surface and flowfield from the lower (windward) surface and flowfield. By exploiting this
flowfield delineation, it may be possible to employ two distinct geometries to improve
performance for boost-glide missions. Ideally, the upper surface can be rounded with a
large radius of curvature to reduce convective heating during pull-up, as found on many re-
entry capsules. Additionally, the lower surface could be a high lift-to-drag ratio waverider
geometry designed to greatly extend the vehicle’s gliding range. These advantages are
particularly significant as the desired range increases. A preliminary exploration of such a
combined capsule/waverider configurations has generated many successful examples.
Particle Swarm Optimizations have been performed to maximize the inviscid lift-to-drag
ratio aerodynamic performance of such concepts. Two-dimensional and three-dimensional
examples have been found and are presented.

Nomenclature
b = semi-span
h = approximate thickness of waverider at fuselage station of the center-of-gravity
n = power-law body exponent
u = streamwise dimension, used in Bezier Surface generation
v = spanwise dimension, used in Bezier Surface generation
CM = pitching moment coefficient
CN = yawing moment coefficient
Cp = pressure coefficient
D = inviscid drag
K = hypersonic similarity parameter (K=2M∞tan(θ))
L = inviscid lift
Lref = reference length used in aerodynamic coefficients
M∞ = freestream Mach number
Sref = reference area used in aerodynamic coefficients
X = streamwise distance measured from the nose of the waverider
Y = spanwise distance measured from the centerline of the waverider
Z = vertical distance measured from an origin
α = angle of attack
β = sideslip angle
θ = local flow turning angle, body flap deflection angle

1
Chief Technology Officer, AIAA Associate Fellow

Copyright © by Patrick E. Rodi. Published by the American Institute of Aeronautics and Astronautics, Inc., with permission.
I. Introduction

T HE pursuit of high lift-to-drag ratio (L/D) hypersonic bodies has been underway for decades. By taking
advantage of the hyperbolic nature of the inviscid flowfield, vehicle performance can be greatly increased. One
such example of this approach is the waverider family of vehicles. Waveriders effectively increase the lift
generated from a vehicle moving through the air at high speed by riding the shock wave that the vehicle itself has
created. Various waverider approaches have been developed since the 1950’s1, 2. For axisymmetric shock shapes,
the Conical Waverider Method has been extensively developed3, 4. However, limits to vehicle L/D ratios exist with
the conical approach. In order to further increase the L/D of waveriders at useful lift coefficients, the Osculating
Cones Method was developed. This approach permits a more general definition of the possible shock wave shapes
and a significant improvement in waverider L/D was obtained5. These results have since been confirmed through
analysis and experimentation6,7. A recent evolution of the Osculating Cones Method is the Osculating Flowfield
Method8. Similar to the Osculating Cones Method, the Osculating Flowfield Method uses a series of planes that are
created normal to the local shock wave shape defined at the trailing edge of the waverider geometry. Currently, a
Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on April 4, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2423

series of “power law body”-based flowfields are employed on each osculating plane. These power-law bodies are
reshaped by changing the exponent, a.k.a. the “n-factor,” on each osculating plane. In both methods, the procedure
used to generate the waverider geometry begins by prescribing the upper surface trace and the shock wave trace on
the baseplane of the vehicle. This baseplane is illustrated in Figure 1. The zero radius (i.e. sharp) leading edge of
the waverider is then found by determining the intersection point between each osculating plane and the upper
surface, at the shock angle generated by the local flowfield on that osculating plane. With the leading edge defined,
the waverider lower surface can be created by tracing streamlines from the leading edge point within each osculating
plane. In addition to manipulating both the upper surface trace and the shock wave trace, the cone angle and power
law factor can vary between osculating planes as long as the assumption of negligible cross plane flow is not
violated. Vehicles generated with the Osculating Flowfield Method have demonstrated modestly superior L/D
performance over those created using the Osculating Cones Method9. Historically, in the vast majority of waverider
designs the potential for upper surface performance contributions is largely ignored as defining the surface in the
freestream direction is the norm. However, the idea of improving waverider performance via changing the upper
surface is not new, and has been examined for waverider cruise flight conditions by the author10.
The Boost-Glide Vehicle (BGV) concept has recently drawn a lot attention from across the world11. In response
to numerous development programs and successful flight tests, the United States has responded with the Falcon
Program’s Hypersonic Text Vehicle-2 long range BGV12, as well as various shorter range applications13. Interest
remains high as BGVs receive continued funding support from the U.S. Department of Defense14. While the current
interest level is high, BGVs have been a subject of study as far back as the 1950’s15-18.
While the waverider construction approaches mentioned above can produce vehicles with high aerodynamic
performance, the sharp (or nearly so) leading edges can produce significantly high heating. This is especially true
during the pull-up maneuver of a boost-glide mission. During this portion of the flight, a blunt geometry producing
reduced heating rates would be preferred.
For low heating rates to the upper surface of the waverider configurations, various blunt configurations offer a
practical solution. A number of studies have examined using a high altitude vehicle that would use aerodynamic lift
and/or drag to perform maneuvers to adjust orbital parameters such as plane change, perigee shifts, and true anomaly
changes19. In the early 1990’s NASA conducted a number of studies for a high altitude/high speed vehicle known as
the Aero-Assisted Flight Experiment20-21. The selected configuration for this mission was a high drag geometry as
shown in Figure 2. Additionally, for both re-entry from low Earth orbit and for lunar return trajectories, various
capsule shapes have been designed. The performance and stability of such configurations has been examined since
the dawn of the manned space program22-24, and through the Apollo program25-27. Four such capsules are illustrated
in Figure 3. All of these blunt configurations are the source of inspiration for the shaping of the waverider upper
surface geometry.
The focus of the current work is to modify the upper surface of various waverider geometries to reduce the
heating rates during the pull-up maneuver by introducing bluntness. This geometric change, along with flying the
vehicle in a suitable attitude during pull-up and executing a reorientation maneuver before the glide portion, along
with the high aerodynamic performance of a waverider configuration, offer the potential of a new class of
hypersonic vehicles known as a Capsule/WaveRider (C/WR).

II. Vehicle Requirements


The most fundamental requirements for a C/WR design is to have stable longitudinal trim points for angles-of-
attack (α) that correspond to suitable 1) capsule re-entry flight and, 2) waverider gliding flight orientations.
Additionally, the geometry is required to contain sufficient volume for equipment and have a center-of-gravity
position that is potentially possible to achieve. Furthermore, aerodynamic performance is to be optimized. In this
study, a value that is a combination of the lift-to-drag ratio during capsule flight and during gliding flight has been
used as the cost function.

III. Two-Dimensional Configuration Optimizations

A. Vehicle Design.
In order to explore the potential viability of such a C/WR concept, optimizations in two-dimensions were
performed to confirm the existence of one or more successful configurations. In these optimizations, a rudimentary
representation of a C/WR configuration was used. This representation was comprised of five points which were
linked together by linear sections, creating the Outer Mold Line (OML) of the geometry. Figure 4 shows a two-
dimensional C/WR configuration in a side view representation. The five points that define the outer mold line are:
1) the upper surface leading edge, 2) a point where a change in the slope of the upper surface may occur, 3) the top
Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on April 4, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2423

of the base region, 4) the bottom of the base region, and 5) a point where a change in the slope of the lower surface
may occur. From point five, the geometry returns to point one to close the configuration. The geometry between
points 1-3 represent the capsule aspects of the configuration. A break-point, where the slope may change, in the
upper surface slope is included to permit the geometry to trim at an α suitable for re-entry and to generate a positive
lift-to-drag ratio in such an orientation. The surface between points 4, 5, and 1 represent the waverider aspects of
the configuration. The break-point in the lower surface is primarily used to generate a trailing edge control surface
that can be sized and orientated to provide trim for the vehicle when in gliding flight. Using the above described
geometric model, a series of optimizations were performed to explore the viability of the C/WR concept. A typical
optimization and the corresponding results, are described below.
To demonstrate the potential of the C/WR concept, a series of Particle Swarm Optimizations28-29 (PSO) were
performed. A number of points were fixed, or otherwise limited, in space. The leading edge, point 1, was fixed at
x=1 and z=1. The geometry was assumed to be a unit length. The fixed location components are listed in Table 1.

Table 1. Fixed Point Information for the Two-Dimensional C/WR Geometries.

# Variable Name Description Fixed Value


1 Xpt1 X-direction location of the first body point 1
2 Zpt1 Z-direction location of the first body point 1
3 Zpt2 Z-direction location of the second body point 1
4 Xpt3 X-direction location of the third body point 1
5 Xpt4 X-direction location of the fourth body point 1
6 Zpt5 Z-direction location of the fifth body point (function of xpt5 and θ)

A typical optimization employed seven independent variables, and an example set is listed in Table 2.

Table 2. Independent Variables for the Two-Dimensional C/WR Geometry Optimizations.

# Variable Name Description Min Value Max Value


1 θ Wedge angle of waverider leading edge 10° 15°
2 Xpt2 X-direction location of the second body point 0.4 0.8
3 Zpt3 Z-direction location of the third body point 0.85 1.0
4 Zpt4 Z-direction location of the fourth body point 0.7 0.8
5 Xpt5 X-direction location of the fifth body point 0.7 0.99
6 Xcg X-direction location of the center of gravity 0.4 0.65
7 Zcg Z-direction location of the center of gravity 0.92 0.96

Together these variables define the locations of the five body points used to describe the geometry, the wedge angle
of the waverider, and the center of gravity location. Inviscid aerodynamic data were calculated over a range of
angles-of-attack from -140° to +20°. Zero sideslip was assumed (i.e. β°=0). The aerodynamic lift, drag, and
pitching moment were calculated using Modified Newtonian flow model on all compression surfaces, and a Cp=0
model on all expansion surfaces. (While more sophisticated models could have easily been employed, these were
judged to be sufficient to prove the existence of a successful C/WR concept.)
As mentioned earlier, the cost function was defined as a combination of the L/D at the capsule α and the L/D at
the waverider α. In this optimization a factor of three was multiplied to the absolute value of the L/D at the capsule
α and this product was then added to the L/D at the waverider α. Since the PSO optimizer employed was designed
to minimize functions, the reciprocal of the L/D sum is the final value of the cost function. A penalty function was
imposed against geometries that did not have two stable trim points over the α range of interest. Stability was
quantified as the derivative of the pitching moment with respect to the angle of attack. As long as this derivative
was negative when the pitching moment passed through zero, the stability requirement was considered satisfied.
For each PSO run, a solution population size of 30 members was initiated by random interrogation of the design
space. Once an acceptable initial population was found, this population was advanced through 50 generations with
the PSO scheme. The convergence history of the best member of each generation is shown in Figure 5.
The optimized two-dimensional configuration is shown in Figure 6. The OML is shown by the blue symbols
joined by blue line segments. The center of gravity is illustrated by the red colored star symbol. For this geometry,
the optimized values of the independent variables are tabulated in Table 3.
Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on April 4, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2423

Table 3. Optimized Values for the Independent Variables of the Two-Dimensional C/WR Geometries.

# Variable Name Description Optimized Value


1 θ Wedge angle of waverider leading edge 14.37°
2 Xpt2 X-direction location of the second body point 0.7496
3 Zpt3 Z-direction location of the third body point 0.8665
4 Zpt4 Z-direction location of the fourth body point 0.7759
5 Xpt5 X-direction location of the fifth body point 0.7036
6 Xcg X-direction location of the center of gravity 0.4330
7 Zcg Z-direction location of the center of gravity 0.9460

The pitching moment for the optimized two-dimensional geometry, measured with respect to the center of
gravity, is shown in Figure 7. Note that this configuration has two stable trim points at α=-60.83° and =-2.623°.
These α conditions correspond to the boost attitude and the gliding attitude, respectively.
The inviscid lift-to-drag ratio of the optimized configuration is shown in Figure 8. The lift-to-drag ratios at the
capsule trim point and at the gliding trim point are -0.642 and +4.221, respectively. Since the lift-to-drag ratio is
negative at α=-60.83°, the vehicle will be orientated with 180° of roll during the boost segment. The gliding
segment will be performed with zero roll and at α=-2.623°.

B. Extruded Two-Dimensional Geometry.


The two-dimensional geometry produced by the optimization effort above has been extruded in the third (y- or
crossflow) direction by a unit width to create a representative three-dimensional geometry. This geometry is shown
in the α=-60.83° (roll angle of 180°) boost attitude and the α=-2.623° (roll=0°) glide attitude in Figures 9 and 10,
respectively. In both images, the freestream air direction is in the positive x-direction, as indicated by the coordinate
axis shown.

C. Three Degree of Freedom Trajectories.


A simple three degree of freedom re-entry trajectory code has been written in Matlab, following the
development in Reference 30. This script generates aerodynamic coefficients using the same impact methods used
in the above described two-dimensional optimization. Typical values for size and weight were defined and three
degree-of-freedom (3DOF) trajectories were generated for both the vehicle in the boost attitude and in the glide
attitude. Using the initial conditions of an altitude of 120 km, an inertial velocity of 7.75 m/sec, and a flight path
angle of -2°. In both trajectories, the attitude was held constant during re-entry. The re-entry trajectories for the
attitudes are shown in Figure 11. The boost trajectory (colored in red) is the result of maintaining the boost attitude
during the flight. The vehicle plunges into the atmosphere, rebounds, and then the altitude is continuously
diminishing. The trajectory is short in duration (≈1500 seconds), as is typical for a capsule type of vehicle. In
contrast the glide trajectory begins with a plunge that is deeper into the atmosphere. This plunge is approximately
20,000 meters (67,000 ft.) deeper, with the associated greater heating rates. This is a result of the much lower lift
coefficient generated by the configuration at the glide attitude. It is this initial plunge that is to be avoided, along
with the associated heating to the waverider. After this first plunge, the glide trajectory continues with a significant
number of oscillations in altitude and a long flight time of approximately 8500 seconds.
A rough illustration of the performance of the capsule/waverider configuration can be constructed by combining
these two trajectories. In one such combination, the conditions during the first rebound of the capsule configuration
are used to transition from the boost trajectory to the glide trajectory. This point is illustrated by the horizonal black
line in Figure 11, which corresponds to an altitude of ~82,000 meter (270,000 ft). The dynamic pressure at this time
is 156 Pa (3.26 psf), which is sufficiently low that a Reaction Control System (RCS) could be used to re-orientate
the vehicle from the boost attitude to the glide attitude. From this point on the glide trajectory, the vehicle would
then follow the remainder of the glide trajectory. This composite trajectory is shown in Figure 12.

IV. Three-Dimensional Modeling


The three-dimensional geometry created by extruding the two-dimensional optimized geometry is not a
practical vehicle. The ability to robustly and quickly modify the leeward surface of a given waverider is also
required to optimize the three-dimensional geometries, as part of this study. To that goal, a number of geometry
modification schemes are being evaluated. Examples of a number of such schemes is presented in Figures 13-17.
Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on April 4, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2423

Figure 13 shows an upper/rearward perspective view of a typical waverider configuration, with a planar surface
modification added to the rear quarter of the leeward surface. This approach, known as “trimming method #1,” is
analogous to the geometry representation used in the two-dimensional optimization. While this approach is
effective and easy to program, the resulting geometry is not desirable. The relatively sharp corner created at the
delineation of the original leeward surface and the region of the planar cut will produce locally high heating rates,
especially in the boost attitude. The corner may also produce separation which will negatively impact the
performance at both the boost and the glide attitudes.
Figure 14 shows an example product of the second geometry modification process. Known as “trimming
method #2,” this method replaces the plane used in method #1 with a ruled surface of constant curvature. The
primary advantage is to provide the ability to exploit the non-linearities of the flowfield to the optimizer. Such
flexibility in the geometric modeling may produce superior results. Again, the sharp corners require addressing to
reduce the heating rates in the boost attitude, and to avoid the potential for flow separation.
Trimming method #3 is closer to the design approach taken by the Aero-assisted Flight Experiment designers.
In this approach a spherical region is imposed on the leeward side of the waverider. An example of using this
approach is shown in Figure 15. The spherical region is clearly visible via the surface faceting. The radius of the
sphere and the location of the sphere’s origin are potential independent variables for the optimization process.
While the majority of the leeward surface acreage is smooth, there remains a corner at the delineation of the original
and modified surfaces.
Figure 16 shows an example result from the fourth geometry modification process. Known as “trimming
method #4,” this method replaces the spherical surface used in the previous method with an ellipsoidal surface. This
introduces additional variables for optimization. However, the corner problem remains.
Figure 17 shows an example result from the fourth geometry modification process. Known as “trimming
method #5,” this method replaces the waverider upper surface with a Bezier Surface that would permit the
optimization of an upper surface geometry. The control points of the Bezier Surface would be the independent
variables. This method intrinsically addresses the corner problem by producing smooth geometries. However,
precisely matching the waverider’s original leading edge trace in three dimensions would require a large number of
control points to define the Bezier Surface. The vast majority of these points would be independent variables
requiring optimization. Such a large number of variables would require significant computational resources to
optimize.
An alternative application of Bezier Surfaces is to not simply replace the upper surface with such a surface, but
to use a Bezier Surface to manipulate the original upper surface points. This is known as “trimming method #6”.
For waverider vehicles, the leading edge cleanly delineates windward flow from leeward flow, therefore no reason
exists to justify adjusting the upper surface along conical planes, osculating planes, etc. Consequently, a simple (∆z)
approach is used to shift the upper surface downward (in the -Z direction) as a function of X and Y. In this method a
Bezier Surface is defined in the X,Y plane that is slightly larger than the projection of the waverider vehicle. The
Bezier Surface Z value represents the change in the Z-direction to be performed at a given X,Y location on the
waverider’s upper surface. Such a Bezier Surface can be defined using a modest number of variables. In this study,
a suite of 4x4 controls points are used to define the Bezier Surface that is third order in u and is third order in v.
(Directions u and v are coordinates in the computational domain that correspond to X and Y in the physical domain.)
Since the Bezier Surface is the change in the Z-direction, the boundary conditions for the surface are to have ∆Z
values of zero along the leading edge of the waverider. This ensures that the original leading edge geometry is not
altered. This boundary condition defines seven of the 16 control point locations. Additionally, to ensure that the
local slope of the upper surface just behind the leading edge is not altered, the four control points behind the
upstream most points are also set to ∆Z=0 to define the slope of the Bezier Surface at the leading edge. Also, zero
spanwise slope of the Bezier Surface along the leading edge requires ∆Z=0 for the three control points inboard of the
leading edge points. Symmetry across the centerline dictates that the ∆Z values on the centerline and in the first
control point row outboard of the centerline are identical. Therefore, of the original sixteen Bezier Surface control
points, only two values of ∆Z are independent variables requiring optimization. A typical matrix of the Bezier
Surface control points is shown in Figure 18. The leading edge of the waverider is to the left, and the baseplane is
just ahead of the row column of control points at X=1.47. The four control points that may have ∆Z≠0 are contained
in the two elongated ovals. These four are comprised to two pairs having the same ∆Z value, at indicated by the
colored ovals identifying the point pairs. These are the pairs of points highlighted in red and blue in Figure 18.
During the optimization process, the optimizer provides the two ∆Z values required to complete the definition of the
∆Z Bezier Surface. A perspective of one such control point distribution, illustrating the relative vertical placement,
is shown in Figure 19.
Since the waverider geometry can vary significantly from one candidate configuration to the next, a pair of
Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on April 4, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2423

normalized variables are used to define the Bezier Surface control point ∆Z values. The ∆Z value is expressed as a
percentage of the waverider’s centerline thickness at the vehicle trailing edge. With the control points defined, the
Bezier Surface can be constructed. This is shown in Figure 20. With the Bezier Surface defined, then the points on
the upper surface of the waverider were translated downward by the ∆Z value interpolated from the Bezier Surface
at the points’ original X,Y location. This simple method executes quickly, is very robust, and doesn’t alter the
tessellation of the upper surface point distribution. This method has been very successful in generating a modified
waverider upper surface with all of the geometric constraints required. Therefore, this method has been used in all
subsequent optimizations of three-dimensional configurations.
The Bezier Surface is then converted into an unstructured grid where the Matlab command scatterInterpolate is
used to perform the interpolation. This surface is shown in Figure 21. With the ∆Z information available, then the
points on the upper surface of the waverider were translated downward by the ∆Z value interpolated from the Bezier
Surface at the points’ original X,Y location. An example of a waverider configuration with the upper surface
modified by Method #6 is shown in Figure 22. This simple method executes quickly, is very robust, and doesn’t
alter the tessellation of the upper surface point distribution. This method has been very successful in generating a
modified waverider upper surface with all of the geometric constraints required. Therefore, this method has been
used in all subsequent optimizations of three-dimensional configurations.
To obtain longitudinal trim in the waverider glide orientation, two methods are available to the optimizer. First,
the n-factor can be altered from a convex geometry, through a conical geometry, to a concave geometry on each of
the osculating planes. This can have the effect of increasing or decreasing the pitching moment in the glide attitude,
respectively. However, low n-factors producing a concave geometry have significantly reduced interior volume, as
compared to convex geometries. In order to preserve as much volume as possible, the lower limit of the n-factor
range we set to unity. Alternatively, provisions for a control surface has been added. This control surface is located
on the windward side of the waverider when in the glide orientation and is known as the body flap. The body flap
straddles the waverider centerline. The body flap size and position is limited to prevent alteration of the global
waverider feature of the shock wave lying on the vehicle’s leading edge.
Three variables are used to define the body flap geometry. First, is the body deflection angle. This angle is
defined with respect to the freestream direction. This definition permits both upward and downward deflection of
the local OML. Second, is the streamwise location of the leading edge of the body flap. Due the wide range of
geometries possible, this location is normalized by the waverider’s centerline length in the X-direction, Lref. Lastly,
the width of the body flap is defined a ratio of the half of the body flap width to the semi-span, b, (i.e. half span).
The leading edge point along the waverider centerline, along with the body flap’s deflection angle are used to define
the equation of a plane for the body flap. The waverider lower surface and base plane points that fall within the
streamwise and spanwise footprint of the body are translated in the Z-direction to lie on the plane of the body flap.
The lower/rear perspective view of an unmodified waverider example in shown in Figure 23. The body flap
leading edging point on the centerline of the vehicle is at X/L=0.8835 in this example. This point is illustrated on
the waverider configuration by the red symbol shown in Figure 24. The example waverider has been modified by
adding a body flap as shown in Figure 25. In addition to the body flap leading edge location shown in Figure 23, the
width was Y/b=0.3033 in both the +Y and -Y directions. The body flap angle, measured with respect to the
freestream direction, was 5.193°. Since the flowfield on each of the osculating planes of the waverider had a cone
angle between 9.9-10°, a five degree body flap angle (measured with respect to the freestream direction) represents a
lifting of the lower surface body flap footprint by approximately 5°.
III. Three-Dimensional Configuration Optimizations

A. Vehicle Design.
With the potential of successful C/WR concepts confirmed by the two-dimensional design optimizations, and a
practical geometry modeling scheme for three-dimensional configurations defined, PSO runs were conducted to
define and optimize the designs. For the generation of these configurations, a set of 21 independent variables is used
to define the basic waverider geometry, modify the upper surface, and add a body flap. These 21 variables are
defined in Table 4, below.

Table 4. Independent Variables for the 21-Variable Three-Dimensional C/WR Geometry Optimizations.

# Description Min Value Max Value


1 Effective cone angle on centerline (Y=0.0) 9.9° 10°
Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on April 4, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2423

2 Effective cone angle at mid semi-span (Y=0.5) 9.9° 10°


3 Effective cone angle at wing tip (Y=1.0) 9.9° 10°
4 Power law body n-factor on centerline (Y=0.0) 1.0 1.1
5 Power law body n-factor at mid semi-span (Y=0.5) 1.0 1.1
6 Power law body n-factor at wing tip (Y=1.0) 1.0 1.1
7 Y location of the shock wave Bezier Curve control point #2 0.0 1.0
8 Y location of the shock wave Bezier Curve control point #3 0.0 1.0
9 Z location of the shock wave Bezier Curve control point #3 1.0 1.2
10 Z location of the shock wave Bezier Curve control point #4 1.0 1.2
11 Z location of the upper surface Bezier Curve control point #1 1.2 1.6
12 Y location of the upper surface Bezier Curve control point #2 0.0 1.0
13 Y location of the upper surface Bezier Curve control point #3 0.0 1.0
14 Z position of the upper surface Bezier Curve control point #3 1.2 1.6
15 Normalized Delta-Z (Z/H) for Bezier Surface at Centerline CP #3 0.30 0.70
16 Normalized Delta-Z (Z/H) for Bezier Surface at Centerline CP #4 0.60 0.70
17 Normalized X location (X/L) of the waverider center-of-gravity 0.50 0.60
18 Normalized Z location (Z/H) of the waverider center-of-gravity 0.0 0.6
19 Body flap angle (measured w.r.t. the freestream direction) 0° 6°
20 Normalized X location (X/L) of body flap leading edge 0.7 0.9
21 Normalized Y location (Y/b) of the body flap sides 0.3 0.5

The first three variables define the spanwise distribution of the local cone angle, from the centerline to mid-
span, to the wing tip. The next three variables define the spanwise distribution (centerline to swing tip) of the power
law body shapes used on each osculating plane. Variables 7-10 are used to define the third order Bezier Curve of
the upper surface trace in the base plane of the waverider. Variables 11-14 define the third order Bezier Curve of
the shock wave trace in the base plane. Within the current tool suite used to generate the Osculating Flowfield
Waverider, the wing tip is located at the point (Y=1, Z=1). The upper surface and shock wave traces are defined by
a pair of third order Bezier Curves. Such curves require four control points, which are numbered from one to four
going from the flowfield centerline to the wing tip of unit semi-span, b. Variables 7-14 are the ranges of the values
of the control points that are permitted to move. This approach quickly generates a wide variety of traces using a
small number of independent variables.
As a group, the first 14 variables define the baseline waverider which is then modified by shaping the entire
waverider upper surface and the (possible) addition of a body flap. Variables 15 and 16 control the Z location of the
control points of the Bezier Surface used to modify the waverider upper surface. Variables 17 and 18 define the X
and Z location of the center-of-gravity of the configuration used to define the pitching moment as a function of
angle-of-attack. (The center-of-gravity is assumed for a given configuration by the optimizer. No measure of the
center of the volume, or any other geometry-based measure is included in the center-of-gravity definition.) The final
three variables are used to locate, size, and the deflection of the body flap. Table 4 also contains a list of the
minimum and maximum values of each variable.
While each of these variables are permitted to vary, as directed by the optimizer during the PSO runs, a number
of the variables were limited in range to values that offered specific advantages to improve the design. The range of
the first three variables, defining the cone angle distribution, was limited to 9.9°-10°. This was imposed to generate
waveriders with significant amounts of volume. Since the upper surface reshaping removes volume, the cone angle
range was limited to the vicinity of the maximum permitted value of 10°. The range of variables 4-6, defining the
power law n-factor, was limited to 1.0-1.1. This range produces power law bodies from cones to convex geometries.
This limit was imposed after a number of early PSO runs showed that trimming the waverider in the glide
orientation was a challenge. Very often, the pitching moment around zero angle-of-attack was significantly nose
down (i.e. CM<0). The nose down pitching moment is exacerbated by using n-factors less than one (i.e. a concave
geometries), as such geometries have higher pressure towards the downstream end of the windward surface. This
region of high pressure causes a nose down pitching moment increment. Variables 17-18, defining the X,Z location
of the assumed center-of-gravity, were limited to a range that earlier optimization runs showed higher inviscid lift-
to-drag ratio performance.
With the variables and ranges defined in Table 4, a series of PSO runs were conducted using 25 members of a
given population, comprising a total of 250 generations. The cost function to be optimized was the sum of the
Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on April 4, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2423

absolute value of the L/D ratio in the high angle-of-attack boost orientation and the L/D ratio in the low angle-of-
attack glide orientation. A coefficient of three was applied to the boost orientation. The absolute value is taken as
shown in the two-dimensional optimizations, some configurations in the trimmed boost orientation actually have a
negative L/D ratio at the angle-of-attack. This is not an issue as the vehicle could be rolled 180° about the
freestream velocity vector to provide upward lift during this portion of the flight. The coefficient was added since
the L/D values in the boost configuration are significantly lower than in the glide orientation. This value (3) was
chosen based solely on engineering judgement.
For a given defined geometry, the aerodynamic performance was defined over a wide range of angles-of-attack
to provide L/D and pitching moment data. To provide the inviscid pressure distribution representations on
compression surfaces for a very blunt capsule geometry as well as for a sleek waverider geometry, a pair of pressure
impact methods were employed. For regions of the geometry with a low flow turning angle (e.g. the windward
surface in the waverider orientation) the Van Duke Unified Method is used. For regions on the geometry with
locally high flow turning angles (e.g. the blunt surface in the capsule orientation) Modified Newtonian was used,
with a Cpmax value of 1.83. The transition between models was made based on the hypersonic similarity variable, K,
(K=2M∞tanθ), with K=1.8 being the delineation value. For regions in the shadow of the freestream direction, a
Cp=0 condition was applied. It is from the integration of these pressure distributions that the lift, drag, and pitching
moment values were defined.
With the lift-to-drag ratio and pitching moment distribution defined over a broad range of angles-of-attack for a
candidate C/WR configuration, the pitching moment distribution is examined to find two angles-of-attack with zero
moment and a stable pitching moment variation. These angles-of-attack are required to be separated by at least 40°.
With the two trim angles-of-attack defined, the inviscid L/D values are found. These values are combined to
generate the cost function, as mentioned above. The Matlab PSO routine performs minimization, so the cost
function is the reciprocal of the composite L/D ratio. The PSO has full control of the 21 variables across the design
space to minimize the cost function.
The optimized configuration for the variables and variable ranges defined in Table 4 is shown in Figure 26.
Overall, this vehicle has a rather flat appearance. This is an effect of the range in the Y-direction values of the
control points for the Bezier Curves used to define the upper surface and shock wave traces. The vehicle has a
pronounced curvature to the upper surface, as seen in the perspective views and the side view. A body flap has been
added to improved trim in the glide orientation. This body flap has an angle, measured with respect to the
freestream direction, of 5.19° degrees. The body flap leading edge is at X/L=0.8835 and is Y/b=0.3033 wide. For
the glide orientation and L/D ratio of 6.764 at an Angle-of-attack of -0.387° was found. In the boost orientation, the
L/D ratio was 0.5147 at an angle-of-attack of -113.1°. At both of these angles-of-attack, the vehicle was trimmed
(i.e. no pitching moment) and stable. The assumed center-of-gravity was located at (X=0.875, Z=1.306). This point
is shown in Figure 26 by the red symbol in the Side View image of the waverider geometry. While this location
appears high in the Z-direction a potential challenge to meet, this location is approximately at the middle of the
centerline profile of the body. The positive value of L/D in the boost configuration indicates that this configuration
need not be rolled inverted during that portion of the flight. The optimized three-dimensional Capsule/Waverider
configuration is shown in the boost orientation in Figure 27. Alternatively, the optimized three-dimensional
Capsule/Waverider configuration is shown in the glide orientation in Figure 28. The cost function of this
configuration is 16% superior to that of the best member of the first generation, demonstrating the effectiveness of
the PSO approach to performing optimization of this problem. From this specific optimization run, and earlier runs,
many thousands of waverider configurations were found that successfully met the requirement of having two stable
trim points for boost and glide orientations. The vehicle described above has the best inviscid L/D-based cost
function value.
While not a part of the optimization process, a directional stability evaluation has been performed on the
optimized three-dimensional configuration described above. In both the boost and the glide orientations, the
optimized waverider was found to be directionally stable. A directional stability requirement may be added to future
optimization of three-dimensional Capsule/Waverider configurations.

B. Three Degree of Freedom Trajectories.


The optimized three-dimensional Capsule/Waverider described earlier has been analyzed using the Matlab re-
entry trajectory code in the boost orientation. Using the same initial conditions as in the trajectories of the two-
dimensional geometries of an altitude of 120 km, an initial velocity of 7.75 m/sec, and a flight path angle of -2°, the
altitude time history presented in Figure 29 was generated. The optimized three-dimensional Capsule/Waverider
generated a similar rebound trajectory as observed with the optimized two-dimensional vehicle, as shown in Figure
Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on April 4, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2423

11 and 12. As in the earlier trajectories, there is ample opportunity in the flight to perform the reorientation
maneuver an extremely low levels of dynamic pressure and aerodynamic forces.
Trajectory analysis of the optimized three-dimensional Capsule/Waverider in the glide orientation is currently
underway. Following the conclusion of this effort, integrated trajectories (i.e. those containing both the boost
portion and the glide portion) will be examined. In future work, trajectory performance measures such as the total
range, maximum convective heating, and total endurance will be integrated into the cost function used in the PSO
runs.

VI. Conclusions
Be employing the delineation of the windward and leeward flowfields by waverider geometries, curvature of the
upper surface can be introduced producing a new class of high speed vehicle, the Capsule/Waverider configuration.
A preliminary exploration of such a combined capsule/waverider configurations has generated many successful
examples. Particle Swarm Optimizations have been performed to maximize the inviscid lift-to-drag ratio
aerodynamic performance of such concepts. Two-dimensional and three-dimensional examples have been found
and were presented.
During the execution of the PSO runs designed to optimize two-dimensional Capsule/Waverider configurations,
hundreds of configurations have been defined that met the basic requirements of having two stable trim points, one
for a boost attitude and one for a glide attitude. The PSO approach has successfully optimized two-dimensional
configurations by maximizing the combined inviscid lift-to-drag performance of the configuration in both boost and
glide orientations. Trajectories performed with the optimized configuration illustrate a pop-up feature that can be
used to provide an opportunity to perform the reorientation maneuver.
A number of shaping methods have been examined for the generation of three-dimensional Capsule/Waverider
configurations. A Bezier Surface method has been selected as the required geometric constraints and features can be
incorporated, while requiring a modest number of independent variables for optimization. The Bezier Surface
defines a value of the Z-variation of the upper surface geometry as a function of the local X,Y locations.
The PSO runs to optimize three-dimensional Capsule/Waverider configurations, have produced thousands of
example configurations that met the basic requirements of having two stable trim points, one for a boost attitude and
one for a glide attitude. Additionally, the optimized configuration was found to be directionally stable in both
orientations. The PSO approach has successfully optimized three-dimensional configurations by maximizing the
combined inviscid lift-to-drag performance of the configuration in both boost and glide orientations. Trajectories
performed with the optimized configuration illustrate a pop-up feature that can be used to provide an opportunity to
perform the reorientation maneuver.

References
1
Nonweiler, T.R.F., “Aerodynamic Problems of Manned Space Vehicles,” Journal of the Royal Aeronautical
Society, Vol. 63, pp. 521-528.
2
Eggers, A.J., Ashley, H., Springer, G.S., Bowles, J.V., and Ardema, M.D., “Hypersonic Waverider
Configuration from the 1950’s to the 1990’s,” AIAA Paper 93-0774, 31th Aerospace Sciences Meeting, Reno, NV,
1993.
3
Bowcutt, K.G., Anderson, J.D., and Capriotti, D., “Viscous Optimized Hypersonic Waveriders,” AIAA Paper
87-0272, 24th Aerospace Sciences Meeting, Reno, NV, 1987.
4
Corda, S., and Anderson, J.D., “Viscous Optimized Hypersonic Waveriders Designed from Axisymmetric
Flow fields,” AIAA Paper 88-0369, 26th Aerospace Sciences Meeting, Reno, NV, 1988.
5
Center, K.B., Sobieczky, H. and Dougherty, F.C., “Interactive Design of Hypersonic Waverider Geometries,“
AIAA Paper 91-1697, AIAA 22nd Fluid Dynamics, Plasma Dynamics and Lasers Conference, Honolulu, HI, June
1991.
6
Mill, R.W., Argrow, B.M., Center, K.B., Brauckmann, G.J., and Rhode, M.N., “Experimental Verification of
the Osculating Cones Method for Two Waverider Forebodies At Mach 4 and 6,” AIAA Paper 98-0682, 36th
Aerospace Sciences Meeting, Reno, NV, 1998.
7
Jones, K.D., and Center, K.B., “Waverider Design Methods for Non-Conical Shock Geometries,” AIAA Paper
2002-3204, 3rd Theoretical Fluid Mechanics Meeting, St. Louis, MO, 2002.
8
Rodi, P.E., “The Osculating Flowfield Method of Waverider Geometry Generation,” AIAA Paper 2005-0511,
th
44 AIAA Aerospace Sciences Meeting, Reno, NV, 2005.
9
Rodi, P.E., Genovesi, D., “Engineering-Based Performance Comparisons Between Osculating Cone and
Osculating Flowfield Waveriders,” AIAA Paper 2007-4344, 37th AIAA Fluid Mechanics Conference, Miami, FL,
2007.
10
Rodi, P.E., “On Using Upper Surface Shaping to Improve Waverider Performance,” AIAA Paper 2018-0554,
Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on April 4, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2423

2018 AIAA SciTech Forum, Orlando FL, 2018.


11
Croft, J., “Weapons Delivery Goes Hypersonic,” Aerospace America, May 2004, pp 38-42.
12
Walker, S.H., and Rogers, F., “Falcon Hypersonic Technology Overview,” AIAA Paper 2005-3253, January
2006.
13
Trimble, S.,” Raytheon Tactical Boost Glide Baseline Review Completed”, Aviation Week and Space
Technology, July 19, 2019, https://aviationweek.com/awindefense/raytheon-tactical-boost-glide-baseline-review-
completed, extracted on 7/30/19.
14
Warwick, G., “DARPA Budget Expands Hypersonic Boost-Glide Weapon Demo,”
https://aviationweek.com/awindefense/darpa-budget-expands-hypersonic-boost-glide-weapon-demo, extracted on
7/30/19.
15
Seiff, A., Allen, H.J., “Some Aspects on the Design of Hypersonic Boost-Glide Aircraft” NASA RM A55E26,
August 1955.
16
Paulson, W., and Shanks, R.E., “Investigation of Low-Subsonic Flight Characteristics of a Model of a
Hypersonic Boost-Glide Configuration having a 78° Delta Wing,” NASA TN D-894, May 1961.
17
Graves, E.B., and Carmel, M.M., “Aerodynamic Characteristics of Several Hypersonic Boost-Glide Type
Configurations at Mach Numbers from 2.30 to 4.63,” NASA TM X-1601, September 1968.
18
Repic E.M., Olson G.A., and Milliken R.J. “A Methodology for Boost-Glide Transport Technology
Planning,” NASA CR-2346, February 1974.
19
Waldberg, G.D., “A Survey of Aeroassisted Orbit Transfer,“ Journal of Spacecraft, Vol. 22, No.1, Jan.-Feb.
1985.
20
Nelson, H.F. (editor), Thermal Design of Aeroassisted Orbital Transfer Vehicles, AIAA Progress in
Aeronautics and Astronautics Volume #96, 1985, ISBN (print): 978-0-915928-94-1.
21
Gupta, R.N, Jones, J.J., and Rochelle, W.C., “Stagnation-Point Heat-Transfer Rate Predictions at Aeroassist
Flight Conditions,” NASA TP-3208, September 1992.
22
Bird, J.D., and Reese, D.E., “Stability of Ballistic Reentry Bodies,” NASA 58E02a, August 1958.
23
Faget, M.A., Garland, B.J., and Buglia, J.J., “Preliminary Studies of Manned Satellites Wingless
Configuration: Nonlifting,” NASA 58E07a, August 1958.
24
Penland, J.A., and Armstrong, W.O., “Static Longitudinal Aerodynamic Characteristics of Several Wing and
Blunt-Body Shapes Applicable for use as Reentry Configurations at a Number of 6.8 and Angles of Attack up to
90°,” NASA TM X-65, October 1959.
25
Boylan, D. E., and Griffith, B. J., “Simulation of the Apollo Command Module Aerodynamics at Re Entry
Altitudes,” Proceedings of the Third National Conference on Aerospace Meteorology, American Meteorological
Society, Boston, May, 1968.
26
Moseley, W.C., Graham, R.E., and Hughes, J.E., “Aerodynamic Stability Characteristics of the Apollo
Command Module,” NASA TN D-4688, August 1968.
27
Crowder R., S., and Moote, J.D., “Apollo Entry Aerodynamics,” AIAA Journal of Spacecraft, Vol. 6, No. 3,
pp. 302-307.
28
Kennedy, J., and Eberhart, R., “Particle Swarm Optimization,” in Proceedings of IEEE International
Conference on Neural Networks, pp. 1942-1948, 1995.
29
Sam, “Another Particle Swarm Toolbox,” Mathworks User Community,
http://www.mathworks.com/matlabcentral/fileexchange/25986-another-particle-swarm-toolbox.
30
Miele, A., Zhao, Z.G., and Lee, W.Y, “Optimal Trajectories for the Aeroassisted Flight Experiment, Part 1,
Equations of Motion in an Earth-Fixed System,” NASA CR-186134, Rice University, 1988.

Figures
Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on April 4, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2423

Figure 1. Baseplane of the Osculating Flowfield Waverider’s Flowfield.

Figure 2. NASA’s Aero-Assisted Flight Experiment Design.


Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on April 4, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2423

Figure 3. Various Capsule Designs.

Figure 4. Two-Dimensional Geometry Representation.


Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on April 4, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2423

Figure 5. Convergence History for the Two-Dimensional Optimization Run.

Figure 6. Optimized Two-Dimensional Geometry.


Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on April 4, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2423

Figure 7. Pitching Moment About the Center of Gravity Across the Angle-of-Attack Range.

Figure 8. Inviscid Lift-to-Drag Ratio Across the Angle-of-Attack Range.


Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on April 4, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2423

Figure 9. Optimized Two-Dimensional Capsule/Waverider Configuration in the Boost Attitude.

Figure 10. Optimized Two-Dimensional Capsule/Waverider Configuration in the Glide Attitude.


Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on April 4, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2423

Figure 11. Capsule/Waverider Boost Attitude vs. Glide Attitude Trajectory Comparison.

Figure 12. Capsule/Waverider Representative Composite Trajectory.


Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on April 4, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2423

Figure 13. An Example of Method #1 – Planar Trimming.

Figure 14. An Example of Method #2 – Cylindrical Trimming.


Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on April 4, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2423

Figure 15. An Example of Method #3 – Spherical Trimming.

Figure 16. An Example of Method #4 – Ellipsoidal Trimming.


Figure 17. An Example of Method #5 – Bezier Surface Replacement.
Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on April 4, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2423

Figure 18. The Top View of the +Y Direction Showing a Typical Control Point Distribution.
Figure 19. The Control Points in Perspective view, Showing the f(u,v) Value.
Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on April 4, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2423

Figure 20. The Control Points and Resulting Bezier Surface.

Figure 21. The Unstructured Representation of the ∆Z Bezier Surface.


Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on April 4, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2423

Figure 22. The Upper/Side Perspective View of a Modified Waverider upper Surface using Method #6.

Figure 23. An Example Waverider Before Body Flap Modification.


Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on April 4, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2423

Figure 24. Location of the Body Flap Point on the Waverider Centerline.

Figure 25. Example Waverider After Body Flap Modification.


Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on April 4, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2423

Figure 26. Optimized Three-Dimensional Capsule/Waverider Configuration.

Figure 27. Optimized Three-Dimensional Capsule/Waverider Configuration in the Boost Attitude.


Downloaded by UNIVERSITY OF TEXAS AT AUSTIN on April 4, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-2423

Figure 28. Optimized Three-Dimensional Capsule/Waverider Configuration in the Glide Attitude.

Figure 29. The Altitude Time History of a Representative Re-entry Trajectory,


for the Optimized Three-Dimensional Capsule/Waverider in the Boost Attitude.

Das könnte Ihnen auch gefallen