Вы находитесь на странице: 1из 8

(c)2000 American Institute of Aeronautics & Astronautics or Published with Permission of Author(s) and/or Author(s)' Sponsoring Organization.

AOO-39726
AIAA-2000-4116
AN ANALYTIC AEROCAPTURE GUIDANCE ALGORITHM
FOR THE MARS SAMPLE RETURN ORBITER

Downloaded by EMBRY-RIDDLE AERO UNIV. on March 26, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2000-4116

James P. Masciarelli*
NASA Johnson Space Center
Houston, TX
and
Stephane ROUSSEAU*, Hubert FRAYSSEf, Etienne PEROT1
Centre National d'Etudes Spatiales
Toulouse, France
Abstract
The Mars Sample Return mission calls for the first use
of an aerocapture trajectory to insert a vehicle into orbit
about Mars. A joint CNES-NASA effort is underway
to study this very critical phase of the mission and
choose the best guidance algorithm to be used for this
atmospheric trajectory. This paper discusses an analytic predictor corrector guidance algorithm, being
jointly developed by CNES and NASA. The algorithm
is being considered for the Mars Sample Return Orbiter
vehicle, which will be developed, launched, and operated by France. Nominal and Monte Carlo trajectory
results are presented for two different versions of the
algorithm. These results show that the algorithm is a
good candidate for the given mission.
Nomenclature
AFE
APC
CD
CL
CNES
D
EMCD
E,0,
g
Gd
G^

Aeroassist Flight Experiment


Analytic Predictor Corrector
Aerodynamic drag coefficient
Aerodynamic lift coefficient
Centre National d'Etudes Spatiales
Deceleration due to drag
European Martian Climate Database
Total orbital energy
Gravitational acceleration
Drag error gain
Altitude rate error gain

Altitude rate

Reference altitude rate


hs
Atmosphere scale height
Orbit inclination
/
JSC
Johnson Space Center
K
Equilibrium glide factor
m
Vehicle mass
Mars GRAM Mars Global Reference Atmosphere
Model
MSRO
Mars Sample Return Orbiter
H
Planet gravitational constant
NASA
National Aeronautics and Space
Administration
0cmd
Bank angle command
q
Dynamic pressure
R
Distance from center of planet
Ra
Apoapsis radius
Atmospheric exit radius
Rod,
Desired velocity at atmospheric exit
Predicted velocity at atmospheric exit
Inertial velocity
Velocity error ( Vai, -Vdes,Kd)
Introduction
The Mars Sample Return mission is a joint CNES and
NASA mission of the Mars Exploration Program1. The
mission requires an orbiter vehicle that will enter Mars
orbit, locate and retrieve the samples, and return them
to Earth. This Mars Sample Return Orbiter (MSRO),
which is scheduled for launch in 2005, will be developed, launched, and operated by France.

* Engineer, Aeroscience and Flight Mechanics Division, Member AIAA


Engineer, Mission Analysis Department
* Engineer, COFRAMI, Toulouse, France
Copyright 2000 by the American Institute of Aeronautics and Astronautics, Inc. No copyright is asserted in the
United States under Title 17, U.S. Code. The U.S. Government has a royalty-free license to exercise all rights under
the copyright claimed herein for Governmental Purposes. All other rights are reserved by the copyright owner.
+

525

Downloaded by EMBRY-RIDDLE AERO UNIV. on March 26, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2000-4116

(c)2000 American Institute of Aeronautics & Astronautics or Published with Permission of Author(s) and/or Author(s)1 Sponsoring Organization.

The MSRO mass is limited to 2700 kg on the launch


pad. In order to maximize the amount of useful mass
delivered to Mars orbit within this limit, the MSRO will
use a guided aerocapture trajectory to insert itself into a
1400 km by 250 km altitude orbit with 45 degree inclination. Over the past year, a joint CNES-NASA team
has been working to study this very critical phase of the
mission, demonstrate its robustness, and choose the best
guidance algorithm to be used for this atmospheric trajectory. A group of atmospheric trajectory specialists
from CNES, NASA-Langley and NASA-JSC has been
working to propose and test guidance algorithms, cross
validate trajectory simulators, and analyze the validation methodology and assumptions for the aerocapture
trajectory. Five guidance algorithms (3 from NASA
and 2 from CNES) are being evaluated. Two of the algorithms are analytical predictor-corrector (APC) algorithms derived from the guidance algorithm developed
for the Aeroassist Flight Experiment (AFE) program2.
These two versions of the APC algorithm were developed independently, one by CNES and one by JSC3.

This paper investigates the use of the APC aerocapture


guidance algorithms for the MSRO vehicle and mission
requirements. First, an overview of the APC guidance
algorithm is given, and the differences between the
CNES derived and the JSC derived versions are discussed. Nominal and Monte-Carlo trajectory simulation results using the two guidance algorithms are then
presented for the expected orbiter arrival conditions at
Mars. These results show that the guidance performance satisfies the MSRO mission requirements and that
the two approaches lead to very similar results. Finally,
the direction of future work on this algorithm is discussed.
APC Guidance Algorithm Overview

The APC guidance algorithm consists of two phases.


The first phase, or capture phase, provides bank commands to stabilize the trajectory and drive the vehicle
toward equilibrium glide conditions. When the vehicle
decelerates to a specified velocity, the second phase, or
exit phase, begins. The exit phase analytically predicts
the velocity vector at atmospheric exit altitude, then
adjusts the lift vector magnitude (through bank commands) so that the velocity achieved at exit altitude will
produce an orbit with the target apoapsis. A significant
feature of the APC algorithm is that no reference trajectories are computed prior to flight; all references are
computed and updated during flight.

oped modified versions of the original guidance. The


changes that were made are discussed below.

Capture Phase

For both the CNES and JSC versions of the guidance,


the capture phase is the same as in [2] except that the
dynamic pressure term in the control equation is replaced by a drag acceleration term. With this change,
the control equation becomes
/

s-i

(*-*/
J

where $cmd is the commanded bank angle, h is the


altitude rate, q is the dynamic pressure, and D is the
deceleration due to drag. Gh and Gd are gains selected
to provide the controller with a low response time
(-100 sec), minimum overshoot of the drag reference,
and a smooth bank angle profile.
During the capture phase, when equilibrium glide is
targeted, the reference altitude rate,/zrey, is zero, and
the reference drag deceleration is given by

where V, is current inertial velocity, R is current radius


vector magnitude, g is acceleration due to gravity, CL
and CD are lift and drag coefficients, and AT is a factor to
determine how much of the lift vector should be used to
maintain equilibrium glide. Since (g-V?IK) is always
negative, the bank angle must be between 90 and
180 deg hi order to maintain equilibrium glide. To provide the controller with some robustness margin, K is
selected to be 1.5, producing a mean value for equilibrium glide bank angle near 135 deg.
Exit Phase

In the exit phase of the guidance, the drag acceleration


term is omitted from the control equation, and the reference altitude rate is chosen to be a constant such that
the velocity at atmospheric exit will provide the desired
target apoapsis altitude. This results in the following
control equation, which is identical to the original version outlined in [2]:

The original version of the APC algorithm was developed for the AFE program, and its derivation can be
found in [2]. CNES and JSC both independently devel-

h-h<ref
= COS AM,. -

526

(c)2000 American Institute of Aeronautics & Astronautics or Published with Permission of Author(s) and/or Author(s)1 Sponsoring Organization.

During the exit phase, Gh is tuned to provide the controller with a response time of 20 to 30 seconds.

With these assumptions, a simplified sequence of calculations is obtained to compute Fra;, and Vdesired. The
details of this derivation can be found in [3]. Figure 1
shows the differences between the original, CNES, and
JSC sequence of computations for the exit phase.

In the original version of the guidance, the reference


altitude rate was chosen by a time consuming iterative
procedure. JSC and CNES developed the following
equation to adjust the reference altitude rate based on
the difference between the predicted and desired exit
velocity instead of using the iterative procedure:

In all of the sequences above, it is necessary to have an


estimate of the current atmospheric density. Density
can be derived from measured drag deceleration, assuming a nominal CD:

desired

ref

Downloaded by EMBRY-RIDDLE AERO UNIV. on March 26, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2000-4116

Estimation of Atmosphere Density

(dVmissldh)

2mDn

where F^, is the predicted velocity at exit, Fdw/red is


velocity at exit to achieve the apoapsis target with given
href, and Vmiss = Fra/, - Vdesired. The difference between
the CNES and JSC versions of the guidance is in how
Vexit and Visited are computed. The CNES version follows a method similar to that in [2], which assumes an
exponential atmosphere and constant radial velocity to
atmosphere exit for estimating velocity loss due to drag.
The JSC version uses these assumptions and the following additional assumptions:
Velocity loss due to drag is instantaneous and can
be applied at any point in the exit phase to estimate
apoapse altitude.
Desired velocity (a square root function of the vis
viva integral) can be approximated by a quadratic
expression and can be calculated at the current altitude rather than the exit altitude.
Original Sequence

SCDV;

The guidance models the density as a simple exponential law of altitude


h-h0
Pmod = Poe

where p0 and h0 are reference density and altitude, and


the scale height, hs, is a constant. The scale height is
selected to provide the best compromised performance
with the anticipated different atmospheric profiles.
A density scale multiplier Kp is defined as the ratio
between the expected density (from the guidance's exponential model) and the measured density. A first or-

CNES Sequence

AF=-

K =

Pmes =

JSC Sequence

V1
AF = -

il

an

h
"exit = n"*

r~ -

Seal

P
Seal

1m

hrefm
CDSqesths

Vr

Fn =

-1
"factor

2FJ4-1
'P\ R2

vdesired"

* yv.

; 2

desirecT~

desired

AF

wish

Figure 1. Comparison of Exit Phase Computation Sequences

527

(c)2000 American Institute of Aeronautics & Astronautics or Published with Permission of Author(s) and/or Author(s)' Sponsoring Organization.

der low-pass filter is used to smooth high frequency


atmosphere disturbances:

Pmod

APC Performance for the MSRO Mission

The density scale multiplier is initialized to 1. A typical value of 0.2 is used for the filter gain, K.

MSRO Vehicle Definition

Note that with this approach, the density scale multiplier will compensate for both density variations and CD
dispersions.

The MSRO vehicle has a mass of 2200 kg at entry. The


vehicle's aeroshell has a reference area of 8.647 m2.
The nominal angle of attack is 2 deg, and for continuum
flow, the nominal values for lift and drag coefficients
are 0.426 and 1.723 respectively. With these characteristics, the vehicle has a nominal lift-to-drag ratio of
0.25, and a ballistic coefficient of 148 kg/m2. The vehicle's attitude control system is capable of producing
bank accelerations of 6.8 deg/s2, and the roll rate is
limited to 20 deg/s. These assumptions correspond to
the CNES MSRO phase A mission design.

Lateral Logic

Aerocapture Mission Requirements

For the MSRO mission, orbit inclination is specified,


but node location is unspecified. Therefore, bank reversals are performed to target only for the desired orbit
inclination. The CNES and JSC versions of the guidance initiate a bank reversal whenever the current inclination exceeds some limit. For the results presented in
this paper, both algorithms used the same lateral logic.

Mission requirements dictate that the post-aerocapture


apoapsis altitude must be 1400 100 km, that the periapse altitude must be greater than -50 km, and that the
orbit inclination must be 45 0.5 deg (Phase A hypotheses). The total AV allocated for post-aerocapture
is 100 m/s. In addition, the aerocapture trajectory must
satisfy the following constraints on deceleration, heat
flux, heat load, and dynamic pressure:
maximum deceleration of 2.8 g's (27.46 m/s2),
maximum heat flux of 460 kW/m2,
maximum total heat load of 70 MJ/m2,
maximum dynamic pressure 4.6 kPa.

Now, the density estimate, dynamic pressure estimate,


and drag deceleration estimate used in the controller
equations are computed as:
1

Pest =Kppmod, qest = pestVr , Dest =


Downloaded by EMBRY-RIDDLE AERO UNIV. on March 26, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2000-4116

manded. Finally, the angular distance to roll is examined. If the distance to roll through lift down is less
than 245 deg, then roll through lift down is commanded, otherwise a roll through lift up is commanded.

qesl.

The lateral logic uses an inclination deadband that is a


function of inertial velocity. Whenever the inclination
error exceeds this deadband, a roll reversal is commanded. In this case, the allowable inclination error,
iaiimabie, is given by
allowable

+
a(v, - vVminj,\iimax ;j11
-t- U\Y/

Mars Arrival Conditions

where V, is the current inertial velocity, a is the change


of allowable inclination error with velocity, /,, and imax
are values that specify the minimum and maximum inclination limits, and Vmin is the velocity where imin is applied.

The roll direction is selected through a series of tests


that examine current velocity, angular distance to roll,
and difference between reference and navigated altitude
rate. First, if the inertial velocity is greater than
5.8 km/s, the direction is commanded based on the
shortest distance to roll. If that condition is not met,
then if the current altitude rate is greater than the reference altitude rate, a roll lift-down is commanded. The
next test computes an allowable error between the reference altitude rate and the current altitude rate. If the
current altitude rate is a certain amount below the reference altitude rate, then a roll through lift up is corn-

528

The arrival conditions corresponding to launch at the


close of launch window were estimated to impose the
most demanding requirements on the aerocapture guidance. Therefore, these conditions were examined first
and are presented in this paper. A launch at the close of
launch window gives an arrival time at Mars of August.
27, 2006 10:28:23 UT. With entry interface defined to
be at a radial distance of 3522.2 km from the center of
Mars, the entry velocity is 5.9 km/s and the flight path
angle is -10.5 deg. The arrival conditions corresponding to launch at the open of the launch window were
later investigated with similar results.
Simulation Overview
The CNES and JSC versions of the APC algorithm
were implemented as subroutines within a 3-degree of
freedom trajectory simulation program. Inputs to the
guidance subroutine include current time, position, ve-

(c)2000 American Institute of Aeronautics & Astronautics or Published with Permission of Author(s) and/or Author(s)1 Sponsoring Organization.

locity, sensed acceleration, and estimated bank angle.


The subroutine outputs commanded bank angle and the
direction to roll. Both versions of the APC algorithm
require on the order of 150 executable lines of
FORTRAN code for the guidance subroutine

Downloaded by EMBRY-RIDDLE AERO UNIV. on March 26, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2000-4116

The guidance algorithms were tested with two different


Mars atmosphere models: the Mars GRAM, provided
by NASA, and the ESA European Martian Climate
Database (EMCD) atmosphere model, provided by
CNES. The vehicle aerodynamic coefficients are determined from tabular data of Q. and CD versus Knudsen number and angle of attack, which was generated
from computational fluid dynamics analysis and wind
tunnel tests of the aeroshell configuration.

shown in Figure 2. Target conditions and trajectory


constraints are satisfied in all cases.

Monte Carlo Results


Monte Carlo simulations were performed to further test
the guidance algorithms. Each run randomly varied the
actual entry position and velocity, the navigated position and velocity, trim angle of attack, aerodynamic coefficients, vehicle mass, and used a perturbed atmosphere density profiles generated with the atmosphere
models. A total of 4000 runs were completed for each
version of the guidance, which includes 2000 cases
with the Mars GRAM, 1000 cases with the EMCD
dusty atmosphere model, and 1000 cases with the
EMCD clear atmosphere model.

Nominal Trajectory Results


Nominal trajectory results obtained with both versions
of the guidance for all of the different atmosphere models look very similar. As an example, nominal trajectory results obtained with the CNES version of the
guidance and the EMCD dusty atmosphere model are

Monte Carlo results obtained with the CNES and JSC


versions of the guidance are shown in Figures. A
summary of the Monte Carlo statistics for the CNES
and JSC versions of the guidance is shown in Table 1.
Again, the two versions of the guidance produce similar
results.

200

D>

500

100

CO

CD

T3

&
CD

CD

CD
-*

-500

-100

-200
3500

4000

4500

5000

5500

-1000

-1500
3500

6000

4000

Inertia! velocity, m/s

4500

5000

5500

6000

5500

6000

Inertial velocity, m/s

46.5

40

46

42
30
c"

1*45.5

I45

CD
O
<B
Q

44
43.5
3500

4000

4500

5000

5500

6000

3500

Inertial velocity, m/s

4000

4500

5000

Inertial velocity, m/s

Figure 2. Nominal Trajectory for CNES Version

529

(c)2000 American Institute of Aeronautics & Astronautics or Published with Permission of Author(s) and/or Author(s)' Sponsoring Organization.

CNES Version
50

65

40

J2 60

530

*55
c

in

JO

Q.

^50

10'
Downloaded by EMBRY-RIDDLE AERO UNIV. on March 26, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2000-4116

1300

1350

1400

1450

45
44.5

1500

Apoapsis Altitude, km

45

45.5

Orbit Inclination, deg

JSC Version

65

50

I
30

c
a
a.

CO

&
CO

^50

20

10'
1300

1350

1400

1450

1500

45

44.5

Apoapsis Altitude, km

45.5

45

Orbit Inclination, deg

Figures. Monte Carlo Results

Table 1. Monte Carlo Statistics (4000 cases each)

CNES Version
Variable

Apoapsis altitude, km
Periapsis altitude, km
Inclination, deg
Max. acceleration, g
Max. dynamic pressure, kPa
Max. heat flux, kW/m2
Max. heat load, MJ/m2

In-plane AV, m/s


Total AV, m/s

Min

Mean

Max

1332.68 1393.48 1493.36


20.10
35.94
43.81
44.58
44.90
45.28
1.52
2.28
3.09
2.24
3.33
4.37
324.68 404.52 479.97
53.44
47.23
62.12
52.40
46.51
61.96
47.39
60.43
75.12

530

JSC Version
Std. Dev.

Min

Mean

Max

30.12 1338.39 1396.70 1486.23


3.97
11.02
30.55
42.36
0.16
44.59
44.89
45.30
0.15
1.52
2.39
3.28
0.27
2.24
3.44
4.67
19.19 324.68 411.83 488.59
1.62
47.01
52.07
61.96
3.14
52.46
60.89
46.76
5.09

47.63

61.07

74.38

Std. Dev.
22.05
4.13
0.17
0.23
0.35
23.47
1.38
1.92
4.91

(c)2000 American Institute of Aeronautics & Astronautics or Published with Permission of Author(s) and/or Author(s)1 Sponsoring Organization.

formance. The Monte Carlo runs used a 3-sigma value


of 0.4 deg, uncertainty in flight path angle. In order to
have as robust a design as possible, it is desirable for
the guidance to not only handle this case, but to capture
as much of the theoretical entry corridor as possible.

Downloaded by EMBRY-RIDDLE AERO UNIV. on March 26, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2000-4116

The Monte Carlo results show that target conditions


were achieved in all of the cases. In the cases with the
EMCD atmosphere model, all trajectory constraints are
satisfied. With the Mars GRAM however, about 2 percent of the cases violated the maximum deceleration or
heat flux constraints. Further investigation into these
cases revealed that the Mars GRAM produces some
high frequency density perturbations that are not seen in
the EMCD model. These high frequency disturbances
caused the deceleration and heat rate limits to be exceeded for very short periods. More work is needed to
determine if this behavior of the model is correct, and if
so, can the vehicle design accommodate these violations of the constraints for the very short time durations
that they occur.

The steep side of the theoretical entry corridor is defined as the flight path angle where a full lift up trajectory just reaches the apoapsis target altitude. At steeper
flight path angles, the vehicle does not have enough lift
to be able to reach the target apoapsis. The shallow
side of the theoretical entry corridor is found by flying
all lift down trajectories to determine the flight path angle where the vehicle just begins to exceed the apoapsis
target. At any shallower flight path angle, the vehicle
does not have enough lift to capture into the target orbit, and faces the danger of skipping out of the atmosphere.

Entry Corridor Width


Although the nominal and Monte Carlo results show
that bom versions of the guidance perform very well for
the MSRO mission, more work was completed to test
the robustness of the guidance. The uncertainty in entry
flight path angle has a major effect on the guidance per-

The theoretical entry corridor width is dependent on the


atmosphere model. For the EMCD model (considering
both clear and dusty scenarios), the theoretical entry
corridor extends from -11.24 to -10.03 deg in flight

Mars GRAM

je

1600

- 1500

j3

ONES

<! 1400

JSC

CO

8- 1300
1200
-11 .5

-11

-10.5

-10

-95

EMCD Dusty

1600
- 1500

ONES
JSC

< 1400
en

55
g-1300

1200-11.5

-11

-10

-10.5

-95

EMCD Clear

1600

ONES
JSC

1400

1200
-11

-11

-10

-10.5

-9.5

Entry FPA, deg

Figure 4. Sensitivity to Entry Flight Path Angle

531

(c)2000 American Institute of Aeronautics & Astronautics or Published with Permission of Author(s) and/or Author(s)' Sponsoring Organization.

path angle. For the Mars GRAM model, the theoretical


entry corridor is -11.14 to -9.71 deg.

Downloaded by EMBRY-RIDDLE AERO UNIV. on March 26, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2000-4116

The sensitivity of the 2 versions of the guidance to dispersions in flight path angle were examined by taking
the nominal entry velocity magnitude and varying the
flight path angle at entry interface over the theoretical
corridor range. The results showing apoapsis altitude
achieved with guidance as a function of entry flight
path angle are shown in Figure 4. The dashed lines in
the figure show the theoretical corridor. As can be
seen, both versions of the guidance capture most of the
theoretical corridor, but neither version captures the entire theoretical corridor. Most of the loss occurs at the
shallow end of the corridor.
A New and Improved APC Algorithm
A joint CNES and JSC effort is in progress, under
CNES responsibility, to develop a single APC algorithm that combines the best features of the CNES and
JSC versions discussed above. In addition, various
modifications are being investigated to further improve
the algorithm's performance and robustness.
Areas of current focus are on increasing the amount of
theoretical corridor captured by the guidance as well as
increasing guidance robustness to atmospheric dispersions, aerodynamic dispersions, and system failure scenarios. Some of the modifications being investigated
include corrections based on measured lift to drag ratio,
and incorporation of a drag reference term in the exit
phase logic. Gain tuning is another area of being investigated, as finding the "best" values for gains requires a tradeoff between precision and robustness.
This cooperative effort is proving to be fruitful, as preliminary results are showing improvements over the
algorithms discussed above.
Summary and Conclusions
This paper has examined two different versions of an
APC guidance algorithm for the MSRO mission. Both
versions were derived independently by CNES and JSC
from an algorithm originally developed for the AFE
program. JSC and CNES made similar but slightly different modifications to the original algorithm. Both
versions replaced the dynamic pressure reference with a
drag acceleration reference. Both versions also replaced an iterative sequence to determine reference altitude rate in the exit phase with a simpler calculus
based method. The main difference between the two
algorithms is the method used to predict the atmospheric exit conditions.
Both versions of the guidance algorithm were tested in
trajectory simulation programs. Nominal trajectories

532

and Monte Carlo results have been developed using two


different atmosphere models. In addition, performance
sensitivity to large deviations from the nominal entry
flight path angle has been investigated. These results
show that both versions of the APC guidance algorithm
perform well for the MSRO mission. No significant
performance differences have been found between the
two versions of the algorithm (with gains kept at the
same values). Some trajectory constraint violations
were discovered during the Monte Carlo analysis with
the Mars GRAM atmosphere, and effort is underway to
address this issue.
The work completed by the CNES-NASA team over
that past year has shown that the APC algorithm is a
very efficient and simple algorithm, and therefore a
good candidate for the MSRO mission. The common
NASA and CNES work on this guidance algorithm has
been very fruitful. Future cooperative work will be
done to further improve and test this algorithm.
References
[1] Lee, Wayne; D'Amario, Louis; Roncoli, Ralph;
Smith, John: "Mission Design Overview for the
Mars 2003/2005 Sample Return Mission," AAS
99-305, AAS/AIAA Astrodynamics Conference,
Girdwood, Alaska, August 1999.
[2] Cerimele, C.J.; Gamble, J.D.: "A Simplified Guidance Algorithm for Lifting Aeroassist Orbital
Transfer Vehicles," AIAA-85-0348, AIAA 23rd
Aerospace Sciences Meeting, Reno, Nevada, January 1985.
[3] Bryant, L.E.; Tigges, M.A.; Ives, D.G.: "Analytic
Drag Control for Precision Landing and Aerocapture," AIAA-98-4572.

Вам также может понравиться