Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

39th AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit AIAA 2003-4951

20-23 July 2003, Huntsville, Alabama

Aerocapture Simulation and Performance for the Titan Explorer Mission

David W. Way*
Richard W. Powell*
Karl T. Edquist*
James P. Masciarelli†
Brett R. Starr*

Abstract

A systems study for a Titan aerocapture orbiter has been completed. The purpose of this study was to
determine the feasibility and potential benefits of using aerocapture technologies for this destination. The
Titan Explorer design reference mission is a follow-on to the Cassini/Huygens exploration of the Saturnian
system that consists of both a lander and an orbiter. The orbiter uses aerocapture, a form of aeroassist, to
replace an expensive orbit insertion maneuver with a single guided pass through the atmosphere. Key
environmental assumptions addressed in this study include: the uncertainty in atmospheric density and high
frequency atmospheric perturbations, approach navigation delivery errors, and vehicle aerodynamic
uncertainty. The robustness of the system is evaluated through a Monte Carlo simulation. The Program to
Optimize Simulated Trajectories is the basis for the simulation, though several Titan specific models were
developed and implemented including: approach navigation, Titan atmosphere, hypersonic aeroshell
aerodynamics, and aerocapture guidance. A navigation analysis identified the Saturn/Titan ephemeris error
as major contributor to the delivery error. The Monte Carlo analysis verifies that a high-heritage, low L/D,
aeroshell provides sufficient performance at a 6.5 km/s entry velocity using the Hybrid Predictor-corrector
Aerocapture Scheme guidance. The current mission design demonstrates 3-sigma success without
additional margin, assuming current ephemeris errors, and is therefore not dependent on the success of the
Cassini/Huygens mission. However, additional margin above 3-sigma is expected along with the reduced
ephemeris errors in the event of a successful Cassini mission.

HYPAS Hybrid Predictor-corrector


Nomenclature Aerocapture Scheme
IRIS Infrared Interferometer Spectrometer
BOC Beginning of Cassini L/D Lift to Drag Ratio
c.g. Center of Gravity LAURA Langley Aerodynamic Upwind
CAD Computer Aided Design Relaxation Algorithm
CFD Computational Fluid Dynamics MER Mars Exploration Rover
DoF Degree of Freedom MGS Mars Global Surveyor
EDL Entry, Descent, and Landing MPF Mars Pathfinder
EOC End of Cassini MRO Mars Reconnaissance Orbiter
GRAM Global Reference Atmospheric POST Program to Optimize Simulated
Model Trajectories
SEP Solar Electric Propulsion
UVS Ultraviolet Spectrometer
DDOR Delta Differential One-way Ranging
* Aerospace Engineer, NASA Langley Research Center DV Velocity Addition
† Aerospace Engineer, NASA Johnson Space Center s Standard Deviation

1
American Institute of Aeronautics and Astronautics
This material is declared a work of the U.S. Government and is not subject to copyright protection in the United States.
AIAA-2003-4951

Background

Aerocapture description
Aerocapture, a form of aeroassist, is a
propellant-less alternative to the currently
requisite all-propulsive planetary capture. Using
drag to decelerate the vehicle, aerocapture
replaces the expensive orbit insertion maneuver
with a single guided pass through the
atmosphere. To date, aerocapture has not been
demonstrated in flight.
In contrast, aerobraking uses many
passes through the atmosphere to reduce the
Figure 1: Aerocapture into Circular Orbit
period of an elliptical orbit. This reduces, but
does not eliminate, the propulsive capture
requirement. Aerobraking has been used are launched together on a single Delta IV-class
successfully in the Martian atmosphere by Mars launch vehicle in 2010. Figure 2 shows a
Global Surveyor (MGS), and Mars Odyssey, and Computer Aided Design (CAD) model of the
is planned for Mars Reconnaissance Orbiter stack packaged in a 4 m launch fairing. A Solar
(MRO). Electric Propulsion (SEP) module and a single
A nominal drag profile associated with Venus gravity assist provide a 6.25 year
the aerocapture pass is designed to remove all of interplanetary cruise to the Saturn system.2
the hyperbolic excess velocity and enough Both the orbiter and the lander are
additional orbital energy to place the spacecraft initially targeted for a direct entry to Titan.
in an elliptical orbit with the desired apoapsis. Thirty days prior to arrival, the orbiter releases
Because of the larger energy requirements, the lander and executes a divert maneuver to the
aerocapture occurs at altitudes much lower than desired aerocapture approach trajectory.
aerobraking. A guidance system is used to target
the desired exit conditions by reacting to changes The orbiter provides a telecom link for
in the atmosphere. Bank angle modulation is the lander during Entry, Descent, and Landing
used to control the rate of ascent/descent, which (EDL) then completes an aerocapture to the
indirectly affects the drag. The flight path angles desired science orbit (a near-polar 1700 km
required to fly full lift-up and full lift-down form circular orbit). Following aerocapture, the
a theoretical entry corridor. heatshield and backshell are jettisoned, and the
orbiter begins a three-year science mission.
Figure 1 diagrams the sequence of
aerocapture events. At the first apoapsis after the
aerocapture pass, a small propulsive maneuver Study Goals
must be completed to raise the periapsis to the
desired altitude. The periapsis must be raised A systems study for a Titan aerocapture
during the first orbit in order to prevent the orbiter has been completed as part of the NASA
vehicle from re-entering the atmosphere a second In-space Propulsion Program.3 The purpose of
time. Another small propulsive burn is typically this study was to determine the feasibility and
performed at periapsis to clean-up any residuals potential benefits of using aerocapture
in the desired science orbit apoapsis. technologies for this destination.4,5 The products
of this study are a reference mission, baseline
systems definition, and technology requirements
Titan Explorer Mission that may be used by scientists, systems
engineers, technology developers, and mission
The Titan Explorer design reference managers in planning future missions. This
mission is a follow-on to the Cassini/Huygens study provides additional value over previous
exploration of the Saturnian system that consists systems studies because of the higher fidelity of
of both a lander and an orbiter.1 Both spacecraft

2
American Institute of Aeronautics and Astronautics
AIAA-2003-4951

include aerobraking orbiters: MGS, Mars


Odyssey, and MRO (scheduled for launch in
2005); direct lander entries: Mars Pathfinder
(MPF), Genesis, Stardust, Mars 2001 Surveyor
Lander (cancelled), Mars Exploration Rovers
(MER), and Mars Science Laboratory (MSL)
(planned for launch in 2009); and aerocapture
proposals: Mars Surveyor 2001 Orbiter
(cancelled) and Mars Premier Orbiter
(aerocapture option not adopted). The current
Lander simulation leverages this experience in
atmospheric flight and applies it to a new
destination, Saturn’s largest moon, Titan.

Orbiter Simulation Development

SEP Prop Trajectory Simulation


Module
To aid in the systems study activity, a
Solar high fidelity Monte Carlo trajectory simulation
Arrays has been developed to simulate flight through the
Titan atmosphere during aerocapture. This
simulation provides data and statistics used to
quantify mission success probabilities, evaluate
Figure 2: Launch Configuration
candidate guidance algorithms, and provide the
technical feedback required for mission and
aeroshell design (aerodynamic loads, maximum
heat rate, integrated heat loads, orbit
the analyses and environmental models that were circularization fuel, etc.).
employed.
The Program to Optimize Simulated
Key environmental assumptions, central Trajectories (POST) is the basis for this
to successful aerocapture, are addressed in this simulation.6 However, several Titan specific
study. These assumptions include the models were developed and implemented to
uncertainty in atmospheric density, high support this work. These models include:
frequency atmospheric variability, approach approach navigation, Titan atmosphere,
navigation delivery errors, and vehicle hypersonic aeroshell aerodynamics, and
aerodynamic uncertainty. Aerocapture risk is aerocapture guidance. These models are
mitigated by quantifying the atmospheric discussed in more detail.
uncertainty based on all available measurements,
designing the vehicle to provide adequate
aerodynamic control authority, developing a Atmosphere
robust guidance system, and incorporating
sufficient margins. The robustness of the system An engineering-level atmosphere
is evaluated through Monte Carlo simulation. model, denoted Titan-GRAM, was developed for
this study.7 Titan-GRAM is similar to and based
upon the Mars Global Reference Atmosphere
Simulation Heritage Model, Mars-GRAM, which has been used for in
the design and operations support for many Mars
The heritage of the Monte Carlo exploration projects.
simulation used in this study is based upon
previous Langley Research Center work on Titan-GRAM atmospheric density
many diverse planetary missions that involve predictions are based on minimum, nominal, and
phases of atmospheric flight. These missions maximum density vs. altitude profiles predicted
by Yelle et al.8 The Yelle models are based on

3
American Institute of Aeronautics and Astronautics
AIAA-2003-4951

Figure 3: Yelle et al. Titan Atmospheric Density Profiles

observations from Voyager 1 radio science,


Infrared Interferometer-Spectrometer (IRIS), and Aerodynamics
Ultraviolet Spectrometer (UVS). Shown in
Figure 3, the Yelle density profiles include An aerodynamic model for the
density variation due to latitude, season, and reference spacecraft has been developed using
diurnal effects as well as measurement high-fidelity computations. The reference
uncertainty. spacecraft has a 70 deg sphere-cone heatshield,
Within Titan-GRAM, an atmospheric similar to the Viking Mars Lander entry vehicle,
density control parameter, fminmax, is used to and a bi-conic backshell. The configuration is
linearly interpolate between the Yelle profiles. shown in Figure 5.
An fminmax of 1.0 corresponds to the maximum Constant normal and axial force
expected density for a given altitude, while fminmax aerodynamic coefficients are used for the
of -1.0 corresponds to the minimum expected aerocapture pass simulation and are based on
density. A sinusoidal variation of fminmax with Langley Aerodynamic Upwind Relaxation
latitude was implemented to simulate latitudinal Algorithm (LAURA) Computational Fluid
density gradients during an aerocapture pass. Dynamics (CFD) results in the hypersonic
Within the trajectory simulation, fminmax regime. LAURA solves the viscous fluid
is varied as a function of latitude to capture the dynamic equations on a structured grid with
expected latitudinal gradients. A perturbation built-in adaptation.9 Thermal and chemical non-
model, based on gravity wave theory, is also equilibrium models are used to calculate the
included for use in the Monte Carlo analysis with high-temperature flowfield behind the bow
a maximum perturbation (1s) of 10% the mean shock.
density. Figure 4 shows a sample of perturbed The high heritage, L/D = 0.25, aeroshell
density profiles generated by Titan-GRAM. configuration provides 3.5 degrees of theoretical
corridor width at a 6.5 km/sec entry velocity. A
higher entry velocity of 10 km/s results in a 4.7
degree theoretical corridor.

4
American Institute of Aeronautics and Astronautics
AIAA-2003-4951

Figure 4: Sample Titan-GRAM Density Perturbations

A pseudo bank controller was


Guidance developed to mimic the dynamics of a flight
control system in a Three Degree-of-Freedom (3-
A Hybrid Predictor-corrector DoF) simulation. These effects include a control
Aerocapture Scheme (HYPAS) aerocapture system time lag and a finite system response,
guidance algorithm was developed and included limited by a maximum angular acceleration and a
in the simulation.10 The HYPAS algorithm uses maximum angular velocity. The bank angle
an analytic method, based on deceleration due to controller analytically calculates the time
drag and altitude rate error feedback, to predict required, and resulting angular travel necessary,
exit conditions and then adjust the bank angle to complete the maneuver to the commanded
command in order to achieve a target apoapsis attitude. It has been found that including this
altitude and orbit inclination at atmosphere exit. type of controller in a 3-DoF simulation provides
a good approximation to Six Degree-of-Freedom
The HYPAS guidance consists of two (6-DoF) dynamics.
phases: the “capture phase”, in which the
guidance establishes pseudo-equilibrium glide Because the aerocapture spacecraft
conditions; and an “exit phase”, in which exit performs bank reversals to maintain inclination
conditions are predicted, assuming a constant accuracy, and because these reversals could take
altitude rate, and the lift vector is adjusted to null as much as15 seconds to complete, the trajectory
the error between predicted and target apoapsis. simulation must model the effects of an attitude
Figure 6 shows the guidance phases during an controller. These bank reversals force the
aerocapture pass.
Bank reversals maintain inclination Backshell
error within desired limits. All reference values .721 m
Access Opening
are computed and updated during flight. The .762 m x .864 m
HYPAS algorithm was adapted for use at Titan, .635 m
and two sets of guidance initialization
.740 m
parameters were developed: one for the 6.5 km/s
entry, and one for the 10.0 km/s entry. Monte Heatshield 3.75 m
Carlo trajectory simulations were run with this
guidance to determine overall aerocapture
Figure 5: Aeroshell Configuration
performance.

5
American Institute of Aeronautics and Astronautics
AIAA-2003-4951

Figure 6: Phases of HYPAS Aerocapture Guidance

spacecraft off of the optimum flight profile that angle dispersions would degrade by
the guidance is trying to follow. Not including approximately 52% with the use of BOC states.
the error produced during this maneuver would
Since Titan is a moon of Saturn with an
result in overly optimistic conclusions regarding
orbital period of approximately 16 days, the
the vehicle’s targeting ability and the required
mission designer has a wide choice in approach
circularization DV.
velocities for any mission opportunity (Titan’s
velocity could either add or subtract from the
Navigation nominal Saturn approach velocity). Intercepting
Titan at different true anomalies easily tailors the
Initial states were provided by a JPL entry velocity, with only small changes in the
navigation assessment that assumed post-Cassini
ephemeris knowledge and the following data
sources: two-way Doppler and ranging, DDOR,
and optical navigation.11 These assumptions
resulted in a 3s delivery flight path angle
dispersion of ±0.93 deg. Figure 7 shows the
delivery footprint in the B-plane. The dashed
line in this figure is a radius vector to the
nominal aim-point.
The three dominant contributors to this
delivery error were Saturn and Titan
ephemeredes, maneuver execution error, and
optical data measurement error. The current
Cassini mission is expected to improve the
ephemeris errors by a factor of six. However,
this improved navigation is not guaranteed, but
rather contingent upon the successful completion
of the Cassini mission. Therefore, both
Beginning of Cassini (BOC) and End of Cassini
(EOC) states were evaluated. For the purposes
Figure 7: Delivery Error in the B-plane
of this study, it was assumed that the flight path

6
American Institute of Aeronautics and Astronautics
AIAA-2003-4951

incoming hyperbolic approach trajectory. Entry Results


velocities of 6.5 km/s and 10.0 km/s were
considered. A 15% increase in flight path angle
dispersions was levied on the higher entry
Monte Carlo Analysis
velocity.
The Table 1 summarizes the entry flight System performance, risk, and
path angle dispersions assumed in this study robustness are measured by generating statistics
along with the theoretical corridor for an L/D = from Monte Carlo simulations of the Titan
0.25. Further navigational assessment is aerocapture. Many (generally 2000) individual
required to validate these assumptions. aerocapture trajectories are simulated with
random perturbations applied to initial entry
conditions, vehicle aerodynamics, vehicle mass
Table 1: Entry Flight Path Angle Uncertainties properties, and Titan atmospheric conditions.
Entry EOC BOC Theoretical This flight simulation is composed of three main
Velocity Ephemeris Ephemeris Corridor parts: a POST2 trajectory simulation, which
6.5 km/s ± 0.93 deg ± 1.42 deg 3.5 deg integrates all of the models discussed above; an
10 km/s ± 1.07 deg ± 1.63 deg 4.7 deg executive Monte Carlo script, which coordinates
the generation and execution of 16 parallel
simulations; and various supporting scripts for
sampling random distributions, compiling and
formatting output data, evaluating metrics and
statistics, and producing plots and figures. Table
2 lists the uncertainties used in this study.

Table 2: Monte Carlo Uncertainties


Category Variable Nominal ± 3s or min/max Distribution
Initial Conditions
x- position 603.3 km From covariance Correlated
y- position -390.8 km From covariance Correlated
z- position 3502 km From covariance Correlated
x- velocity -3.363 km/s From covariance Correlated
y- velocity -4.123 km/s From covariance Correlated
z- velocity -3.734 km/s From covariance Correlated
Atmosphere
Perturbation seed 1 1/9999 Uniform (integer)
Fminmax bias 0 -0.53/+0.53 Uniform
Aerodynamics
Trim angle-of-attack -16 deg ± 2.0 deg Normal
CA (axial force) 1.48 ± 3% Normal
CN (normal force) -0.05 ± 5% Normal
Mass Properties
Axial c.g.(Zcg/D) 0.1979 ± 0.00848 Normal
Radial c.g.(Xcg/D) 0.0231714 ± 0.00184 Normal

7
American Institute of Aeronautics and Astronautics
AIAA-2003-4951

Beginning of Cassini

The first scenario examined is for a


navigation ephemeris uncertainty consistent with
knowledge prior to the Cassini mission. This
combination of large navigation uncertainty and
low entry velocity, 6.5 km/s, results in the most
challenging conditions. Statistics for apoapsis
altitude and circularization DV are presented in
Table 3.
Figure 8 shows the aerocapture corridor
(flight path angle) as a function of fminmax
(density). The theoretical corridor is bounded by
the full lift-down and full lift-up cases. The plus
(+) symbols show the range of fminmax during the
active guidance portion of the aerocapture pass,
due to latitudinal variation of fminmax. The circles Figure 8: Entry Corridor, BOC 6.5 km/s
indicate the fminmax at periapsis. The bias in the
data towards the higher values of fminmax is again
due to the latitudinal variation of fminmax, since the
aerocapture pass occurs over northern latitudes.
The target flight path angle was chosen to
capture as many of the cases as possible into the
theoretical corridor. The large spread in entry
flight path angle, compared to the theoretical
corridor, suggests small margins.

Table 3 Performance Statistics, BOC 6.5 km/s


Apoapsis Circularization
altitude (km) DV (m/s)
Minimum 1240.5 157.2
Maximum 2166.3 293.3
Mean 1691.3 179.6
1s ± 63.9 ± 11.1
Figure 9: Apoapsis Altitude, BOC 6.5 km/s
3s ± 191.8 ± 33.2

Figure 10 shows the histogram for the


Figure 9 shows a histogram of the final required D V. This D V includes both the
apoapsis altitude. The mean apoapsis of 1691.3 periapsis raise maneuver and the final
km, with a standard deviation of +/- 63.9 km, circularization burn. The 3-sigma (99.86 %-tile)
compares well with the 1700 km target altitude. value is 212.9 m/s.
Additionally, the 3s range of +/- 191.8 km is
within the target range of < 200 km, which
indicates that the guidance can be tuned to End of Cassini
capture a large percentage of the theoretical
corridor. Only one case (#505) failed to capture. These Monte Carlo results are
However, this case represents a 4.5-sigma case representative of navigation ephemeris
for entry flight path angle, which has a uncertainty post-Cassini for a 6.5 km entry
probability of occurrence of only 1 in nearly velocity. Statistics for apoapsis altitude and
15,000. circularization DV are presented in Table 4.

8
American Institute of Aeronautics and Astronautics
AIAA-2003-4951

Figure 10: Circularization DV, BOC 6.5 km/s


Figure 11: Entry Corridor, EOC 6.5 km/s

Figure 11 shows the aerocapture


corridor (flight path angle) as a function of fminmax
(density). The effect of the improved (reduced)
uncertainty expected from the Cassini mission is
evident by the tighter grouping in flight path
angle as compared to Figure 8. The size of the
flight path angle dispersions, compared to the
theoretic corridor, suggests increased margins for
this scenario. Because of the arrival geometry,
the aerocapture pass occurs over northern
latitudes – entering over the northern pole and
exiting near the equator. Therefore, the mean
fminmax (~0.3) is positive. The target flight path
angle was chosen to bring this mean to the center
of the theoretical corridor.

Figure 12: Apoapsis Altitude, EOC 6.5 km/s


Table 4 Performance Statistics, EOC 6.5 km/s
Apoapsis Circularization
altitude (km) DV (m/s)
Figure 12 shows a histogram of the final
Minimum 1327.8 156.0
apoapsis altitude. The mean apoapsis is 1697.7
Maximum 2196.6 252.8 km, with a standard deviation of 63.5 km. The
3-sigma range of +/- 190.4 km is within the
Mean 1697.7 177.7
target range of < 200 km. All cases captured
1s ± 63.5 ± 9.5 within 500 km of the target apoapsis.
3s ± 190.4 ± 28.6 Figure 13 shows the histogram and
statistics for the required D V. The 3-sigma
(99.86 %-tile) value is 206.3 m/s. The maximum
DV was 252.8 m/s.

9
American Institute of Aeronautics and Astronautics
AIAA-2003-4951

traded against increased Thermal Protection


System requirements.
4. The Monte Carlo analysis demonstrates that
the HYPAS guidance is robust and provides
acceptable performance. Approximately
92% of the theoretical corridor is captured
using this algorithm while requiring only
slightly more than 200 m/s of on-orbit DV to
achieve the target science orbit.

References

1. Bailey, R., Hall, J., and Spilker, T.,


“Titan Aerocapture Mission and
Spacecraft Design Overview,” AIAA-
2003-4800, Conference Proceedings of
Figure 13: Circularization DV, EOC 6.5 km/s the 39th AIAA/ASME/SAE/ASEE Joint
Propulsion Conference and Exhibit,
Huntsville, AL, July 2003.
2. Noca, M., Bailey, R., and Dyke, R.,
“Titan Explorer Mission Trades from
the Perspective of Aerocapture,” AIAA-
Conclusions 2003-4801, Conference Proceedings of
the 39th AIAA/ASME/SAE/ASEE Joint
1. The JPL navigation analysis identified the Propulsion Conference and Exhibit,
Saturn/Titan ephemeris error as major Huntsville, AL, July 2003.
contributor the delivery error. The current
3. James, B. and Munk, M., “Aerocapture
mission design demonstrates 3-sigma
Technology Development Within the
success, without additional margin,
NASA In-Space Propulsion Program,”
assuming BOC ephemeris errors, and is
AIAA-2003-4654, Conference
therefore not dependent on the success of
Proceedings of the 3 9th
the Cassini/Huygens mission. However,
AIAA/ASME/SAE/ASEE Joint
additional margin above 3-sigma is expected
Propulsion Conference and Exhibit,
along with the reduced EOC ephemeris
Huntsville, AL, July 2003.
errors in the event of a successful Cassini
mission. 4. Hall, J., Noca, M., and Bailey, R., “Cost
-Benefit Analysis of the Aerocapture
2. Uncertainty in the Titan atmospheric
Mission Set,” AIAA-2003-4658,
density, including high frequency
Conference Proceedings of the 39th
perturbations, is the single largest unknown.
AIAA/ASME/SAE/ASEE Joint
To mitigate this risk, sufficient margin and
Propulsion Conference and Exhibit,
conservatism are carried in the design of the
Huntsville, AL, July 2003.
entry conditions, aeroshell, and guidance
system. While arrival during a particular 5. Lockwood, M. K., “Titan Aerocapture
season would reduce the expected density Systems Analysis,” AIAA-2003-4799,
range, the full density range was used in the Conference Proceedings of the 39th
Monte Carlo analysis. AIAA/ASME/SAE/ASEE Joint
Propulsion Conference and Exhibit,
3. The Monte Carlo analysis verifies that a
Huntsville, AL, July 2003.
high-heritage, low L/D, aeroshell provides
sufficient performance at a 6.5 km/s entry
velocity. A mid L/D aeroshell technology
development is not required. Additional
aerocapture performance is also available at
higher entry velocities, 10 km/s, but must be

10
American Institute of Aeronautics and Astronautics
AIAA-2003-4951

6. Bauer, G. L., Cornick, D. E., and


Stevenson, R., “Capabilities and
Applications of the Program to
Optimize Simulated Trajectories
(POST),” NASA CR-2770, February
1977.
7. Justus, C. and Duvall, A., “Engineering-
Level Model Atmospheres for Titan and
Neptune,” AIAA-2003-4803,
Conference Proceedings of the 39th
AIAA/ASME/SAE/ASEE Joint
Propulsion Conference and Exhibit,
Huntsville, AL, July 2003.
8. Yelle, R. V., Strobel, D. F., Lellouch,
E., and Gautier, D., “Engineering
Models for Titan's Atmosphere,” in
ESA report SP-1177, “Huygens
Science, Payload and Mission,” August
1997.
9. Cheatwood, F. M. and Gnoffo, P. A.,
“User’s Manual for the Langley
Aerodynamic Upwind Relaxation
Algorithm (LAURA),” NASA TM-
4674, April 1996.
10. Masciarelli, J. and Queen, E.,
“Guidance Algorithms for Aerocapture
at Titan,” AIAA-2003-4804,
Conference Proceedings of the 39th
AIAA/ASME/SAE/ASEE Joint
Propulsion Conference and Exhibit,
Huntsville, AL, July 2003.
11. Haw, R., “Approach Navigation for a
Titan Aerocapture Orbiter,” AIAA-
2003-4802, Conference Proceedings of
the 39th AIAA/ASME/SAE/ASEE Joint
Propulsion Conference and Exhibit,
Huntsville, AL, July 2003.

11
American Institute of Aeronautics and Astronautics

You might also like