Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

AAS 00-109

PRECISE FORMATION FLYING CONTROL OF


MULTIPLE SPACECRAFT USING CARRIER-PHASE
DIFFERENTIAL GPS1

Gokhan Inalhan 2 , Franz D. Busse 3


and Jonathan P. How 4

Formation flying is a key technology for deep space and orbital applications
that involve multiple spacecraft operations. Imaging and remote sensing
systems based on radio interferometry and SAR require very precise (sub-
wavelength) aperture knowledge and control for accurate relative data col-
lection and processing. Closely tied with the Orion and TechSat21 projects,
this work describes the ongoing research to investigate precise relative sens-
ing and control via differential GPS for multiple spacecraft formation flying.
Specifically, we present an autonomous control architecture for formation
flying that integrates low-level satellite control algorithms (formation keep-
ing and relative error correction) with high-level fuel/time optimal forma-
tion coordination and planning. The basic features of this architecture are
implemented on a nonlinear orbital simulation of Orion vehicles with dis-
turbances and a realistic differential GPS measurement model. Also, we
generalize closed-form solutions of passive apertures for constellations with
mean formation eccentricity.

INTRODUCTION
Formation flying is a key technology for deep space and orbital applications that involve
multi-spacecraft operations [1, 2, 3]. Imaging and remote sensing systems based on radio
interferometry and SAR involve relative data collection and processing over an aperture [4,
5, 6, 7] where the resolution of the observations is inversely proportional to the baseline
lengths. For this reason, orbital and deep space distributed apertures formed by formation
flying spacecraft provide desirable characteristics compared to earth based or centralized
apertures located on one structure. Also, the ability to easily form and reconfigure very
long baselines for uniform and dense u-v (observing) plane coverage [8, 9] provides a clear
advantage over traditional systems.
However, there are major technical challenges in achieving the aperture coordination,
control, and monitoring of these distributed vehicles that will be necessary to achieve the
stringent payload pointing requirements [6, 7]. In addition, the fleet control design must be
done with careful consideration of the onboard computation, inter-vehicle communication,
and power requirements for the formation flying spacecraft. Thus a systems-level approach is
essential, allowing explicit inclusion of the hardware limitations (power, mass, computation)
in the theoretical analysis of the various multi-level control architectures [14, 16, 19]. For
1
This work is funded in part under Air Force grant # F49620-99-1-0095 and NASA GSFC grant #NAG5-
6233-0005
2
Research Assistant, Dept. of Aeronautics and Astronautics, ginalhan@stanford.edu
3
Research Assistant, Dept. of Aeronautics and Astronautics, teancum@stanford.edu
4
Assistant Professor, Dept. of Aeronautics and Astronautics, howjo@sun-valley.stanford.edu

1
Fig. 1: ORION: On-orbit demonstration of Fig. 2: Schematic of the Orion Satellite
formation flying. Nonlinear simulation with re- (0.45m cube with cold gas propulsion and
alistic disturbances and actuator limitations. GPS sensing)

this reason, the following system level challenges for such distributed systems have been
identified under the overall formation flying problem:
1. Sensing and metrology (relative/absolute sensing, sensor fusion)
2. Aperture optimization (orbit and formation selection)
3. Fleet and vehicle autonomy (control architecture)
4. Control computation (formation planning, maneuvering and data collection)
5. Spacecraft bus design (RCS, crosslink communications and CDH)
This paper describes ongoing research at Stanford University to address these systems-level
issues for two future on-orbit demonstrations of formation flying: Orion and TechSat21 [13,
18]. The particular emphasis of this work is to demonstrate precise relative sensing and
formation flying control of multiple spacecraft via carrier-phase differential GPS.

FORMATION FLYING ON-ORBIT:


ORION & TECHSAT21
TechSat21 is an AFRL/DoD program focused on the development and on-orbit demon-
stration of various formation flying technologies [18]. Orion is a high-risk, low-cost NASA-
funded project dedicated to the research and development of microsatellites that are capable
of performing distributed relative sensing and formation flying. Orion and TechSat21 rep-
resent major stepping-stones/technology demonstrators for many of the basic control and
sensing elements of future formation flying missions. The specific goals of Orion include:
• Ability to organize a group of small satellites into a pre-determined formation on-orbit.
In particular, this will include the ability to exchange GPS, position and scientific
data between the satellites, as well as execute pre-planned, organized maneuvers [17].
The execution of the maneuvers will be governed by on-board real-time autonomous
control software (typical simulation results shown in Fig. 1)

2
• Use and operation of a low-power, low-cost, multi-channel GPS receiver for real-time
attitude and position determination of the satellites (antenna layout in Fig. 2).
• On-orbit autonomy using various control architectures. Demonstrate autonomous
mode switching from coarse to fine (and back) formation flying.
• Ability to perform formation flying and station keeping with various baselines.
With these main objectives for the Orion and TechSat21 projects, we have developed,
simulated, and experimentally demonstrated basic features of an autonomous control archi-
tecture for formation flying. Low-level satellite control algorithms (formation keeping and
relative error correction) are integrated with high-level, operations-driven, fuel/time opti-
mal formation coordination and planning [22] within a multi-layered architecture. Fig. 3
shows the basic elements and the information flow for this control architecture.
The formation planning part of this architecture was implemented on the Formation
Flying Test Bed (FFTB - free flying vehicles on a granite table) to perform a nonlinear “rigid
body” formation retargeting maneuver requiring a coordinated translation and rotation
while keeping the relative distances fixed [21, 22].
This paper provides further analysis of the three key components of the autonomous
control architecture in Fig. 3 and extends it to orbital operations. In particular, we illustrate
the design and analysis of the low-level regulator for relative error correction and formation
keeping, present results from the CDGPS estimator, and discuss the implementation of a
coordinator for passive aperture formation. All of these components are integrated on an
commercial high-fidelity nonlinear orbit propagation tool [31]. The simulations include real-
istic disturbance models, measurement errors, and typical propulsion system nonlinearities
such as finite thrust and minimum impulse bit.
Within the context of this autonomous formation control architecture, the following
sections present detailed discussions of the tools developed to address the system level issues
(e.g. sensing & metrology, vehicle & fleet autonomy) associated with the coordination and
control of the multiple vehicles in the fleet. Specifically, we present the algorithms and
methods used in control input generation for high-level coordination and low-level satellite
control based on CDGPS measurements (absolute and relative).

FORMATION CONTROL ARCHITECTURE


Within the architecture shown in Fig. 3, autonomy is based on intelligent decisions pro-
vided by the High-Level Coordination in two separate levels. The first level involves vehicle
autonomy where system-level decisions, checks, and updates are conducted under the GNC
Housekeeping algorithm. In addition, this process provides an operations and flight status
link between the GNC subsystem and satellite command and data handling unit, which
explicitly controls the satellite bus and hardware activities. The second level includes fleet
autonomy handled by the Multiple Vehicle Formation Coordinator. In a centralized appli-
cation, the coordinator essentially acts as a fuel efficient “distilling” algorithm that uses
information such as the status and location of each spacecraft to determine the desired
reference orbits (mean element update) and desired individual locations for each satellite.
This complex decision process is tied to the fleet objectives and operations. The next sec-
tion discusses tools developed to perform this coordination. Note that the architecture is
sufficiently flexible that distributed coordination could be used for large fleets of vehicles.

3
Fig. 3: Autonomous Control Architecture

As shown, the major hub of the external GNC information flow is handled via the
crosslink unit with two different data rates. The high data rate intersatellite communication
is needed for the raw GPS data used by the relative navigation process that is part of
the differential GPS (CDGPS) unit. The CDGPS unit consists of 3 primary algorithms
that solve for the absolute, relative, and attitude measurement information. The low data
rate intersatellite communication is used to handle the system-level information regarding
formation flight modes, desired reference orbits and locations, and updates of the fleet
parametric models. With high fidelity on-orbit propagators, these estimators provide state
and parametric model estimates of the current satellite and environment models. All of this
information and the flight mode (based on the operations plan) are processed by the high-
level coordinator to determine the actual low-level satellite attitude and position control.
The algorithms that have been developed for the low-level satellite control are discussed in
the following sections.

High-Level Coordination

In this modular architecture, the coordinator explicitly handles the task of coordination
and scheduling of the operations of the formation based on mission defined objectives. The
majority of the mission operations involve task distribution, formation selection/planning

4
Precise Formation Flying in LEO
Orion
1
Orion
2
Orion
3
0.1

0.05
Out−of−track Separation[km]

−0.05

−0.1

−0.15
0.04
0.02
0
−0.02
−0.04 0.3
0.2
−0.06 0.1
−0.08 0
Radial Separation [km] −0.1
−0.1 −0.2
Intrack Separation [km]
Fig. 4: Typical Nonlinear Formation Initialization Simulation With 3 Orion Spacecraft.

[12], collision avoidance and operation mode selection (e.g. coarse formation flying, precise
formation flying, parking orbit) while performing imaging and remote sensing experiments.
How the computational aspects of the formation planning are handled and implemented will
strongly depend on the number of vehicles in the fleet and their computational capabilities.
For a small fleet (typically 3-4 vehicles ), all duties of the coordinator can be implemented
as a centralized process.
For such a case, the multiple vehicle fuel/time-optimal formation planning can be solved
as a convex Linear Programming (LP) optimization problem using the linearized relative
dynamics [21, 22, 28]. Refs. [21, 22] have previously shown that fuel-optimal trajectories [11]
and thrusting sequences of the vehicles can be generated for typical cooperative maneuvers
such as initialization, resizing, and retargeting of passive apertures. Note that this LP design
process used by the coordinator explicitly accounts for the bounded thrust capabilities and
the limited fuel capacity of each vehicle [20].
Fig. 4 shows the results from a typical formation initialization and precise formation
keeping maneuver in a nonlinear simulation with standard LEO disturbances and DGPS-
level measurement errors. In this implementation, two Orion spacecraft are moved from
their initial positions to predefined formation locations with a ±50m intrack separation
with respect to the central spacecraft. The LP algorithm was used by the coordinator
to design the fuel optimal trajectories to move these two vehicles. After the initialization
is complete, the coordinator selected the precise operation mode to keep the formation
within a 0.5×2.5×2.5m radial/intrack/cross-track error box. In this operation mode, linear
quadratic regulators with deadband are used for stationkeeping.
The LP optimization can be used to answer many interesting aspects of optimal coop-

5
Time Optimal Formation Flight via Differential Drag −5
x 10 Drag Panel Normalized Acceleration
2.5
2
Initial Radial Separation : 300m Trajectory
Location at every orbit
Initial Intrack Separation : 750m
2 1.5

1.5 1

1 0.5
x−radial [km]

Drag [m/s ]
2
0.5 0

0 −0.5

−0.5 −1

−1 −1.5

−1.5 −2
−5 0 5 10 15 20 25 30 35 40 0 20 40 60 80 100 120 140 160 180
y−intrack [km] k step

Fig. 5: Time Optimal Formation Flying Fig. 6: Drag Panel Sequence for Coopera-
via Differential Drag tive Control

erative control of multi-vehicle formation flight. For example, although differential distur-
bances can result in excessive fuel usage, such effects can also be used to the advantage of
the formation flight [29]. Fig. 5 shows a cooperative time optimal formation flying control
of two vehicles using drag panels as the actuation. This problem was solved using the LP
technique. With an initial radial separation of 300m and 750m in the intrack direction, the
vehicles operate their drag panels to come into close proximity. Fig. 6 shows the opening
and closing sequence for the drag panels. The positive and negative effect is obtained by
cooperation between the two vehicles.
For close proximity formations on the order of a few hundred meters, the linearized
dynamics provide useful and precise models for formation flight design. However for larger
or longer maneuvers, which can take more than a few orbits, the effects of measurement
noise, nonlinear orbital effects, and differential disturbances will cause deviations in the
final relative states. For these cases, the following standard techniques could be utilized to
handle the errors in the linearized dynamics:
• Iterative Procedure: By using updates on the relative states of the vehicles, the co-
ordinator could iterate and replan the maneuver to account for any large formation
deviations and anomalies encountered during the execution of the initial plan.
• Inclusion Method: Based on initial plan, all of the unaccounted effects can be traced
along the initial trajectory design. These calculated effects could then be included in
the replan to capture the approximate magnitude of the neglected effects. Note that
a similar idea is used in the low-level satellite control subsection to design a long-term
formation keeping trajectory with differential disturbances (see Fig. 10).
• Trajectory Morphing: Ref. [27] presents an approach that explores the trajectory
space of a nonlinear system by starting from simplified models. With the addition of
linear time-varying feedback, feedforward, and homotopy, the linear trajectories can
then be morphed into nonlinear ones, with the result that stable trajectory tracking
is obtained.
• Nonlinear Optimization: Can also use the linearized trajectory to initialize a nonlinear
optimization technique. A similar approach was successfully implemented on the
FFTB for the experimental nonlinear “rigid-body” retargeting maneuver [22].

6
FORMATION INITIALIZATION OF 8 SPACECRAFT
SC1
SC2
300 SC3
SC4
SC5
200
SC6
SC7
100 SC8
z−outofplane [m]

−100

−200

−300

−400
400

200

−200 −600
−400
−200
0
200
−400 400
x−radial [m] 600
y−intrack [m]

Fig. 7: Formation Initialization Maneuver Selected by the Coordinator

For fleets with a large number of vehicles, the computational aspects of the formation
planning can become very difficult given the amount of information flow and the processing
required. In such a case, the formation coordination can be distributed over the whole fleet
with a distributed coordination scheme which involves an individual bidding process [12].
Fig. 7 shows the result of such an implementation where eight vehicles initialize to a closed-
form ellipse (e = 0) after a bidding and selection process for their locations. Initially all of
the vehicles are on a intrack formation separated by 100m from each other. The only hard
constraint on the planning process is that the vehicles should be placed with equal phasing
along a closed-form ellipse with a semi-major axis of 600m.
Given each vehicles fuel state and initial location, the main issue is to determine which
vehicle should move to each location on the target ellipse. The key parts of the distributed
solution to this problem are the bidding and selection process carried out by each of the
satellites (shown in Fig. 8). Each satellite analyzes the alternative final locations (over a
discrete grid of 1◦ resolution) and associates a cost with each of them. Fig. 9 shows a typical
fuel usage vs. aperture location (phase angle) map created by one of the spacecraft. The ∆V
calculations of the individual spacecraft are solved using the standard LP method discussed
previously. These calculations are very simple because they involve only the vehicle itself.
For this example, the individual bidding decisions of the satellites are based on their
fuel usage vs. aperture location map. With all the bids obtained, the coordinator starts

7
DV MAP vs APERTURE LOCATION FOR SC # 1
0.95

0.9

0.85

0.8

Dv Usage [m/s]
0.75

0.7

0.65

0.6

0.55

0.5
0 50 100 150 200 250 300 350 400
Aperture Phasing [deg]

Fig. 8: Bidding Process Organized by the Coor- Fig. 9: Typical Fuel Usage vs Aperture
dinator Location Map Generated by Each Space-
craft for Bidding Process

identifying possible optimal aperture locations with the highest bidder for any starting
aperture location. Having set that location and the vehicle, all other possible aperture
locations are filled one-by-one using the best remaining highest bidders. This same process
is run again for every phase angle over the discrete grid of 1◦ resolution. After this step,
the coordinator selects the best 3 out of the 360 possible configurations (each consists of
all 8 vehicles with equal phasing around the ellipse). Collision avoidance checks are then
run on these 3 cases to decide on the best final configuration. The selection and collision
avoidance process is not computationally intensive and can be run on any single vehicle (or
be distributed). Note that it would also be possible to include rebidding and learning for
each vehicle under an iterative scheme that uses knowledge of the previous bids made and
the trajectories obtained (would be used to modify the spacecraft bidding strategy).
Although this process is not guaranteed to be globally optimal, it provide some key
benefits because it 1) distributes the computational intensity, 2) removes the possible threat
of single point failure, and 3) allows the vehicles to develop individual decision models
that will be reflected in their bids. For example, a vehicle can decide on a maneuver
which requires minimum amount of orientation requirements due to a reaction wheel failure.
Although the selected location in the aperture might be not the best fuel efficient choice, the
bidding process allows the individual vehicles to include such complex decision parameters,
which is very difficult to incorporate in a generalized optimization due to discrete logical
nature and case dependence.

Low-Level Satellite Control

The low-level satellite control combines linear-quadratic regulators with a formation keeping
algorithm [15]. The linear-quadratic regulators are stored in a gain-schedule table based on
different operational characteristics that are determined by the control effort and relative
error weights. These regulators are used for trajectory tracking and relative error correction
to place the spacecraft in the desired relative error boxes.
Once the spacecraft are placed within the desired error boxes, the controller switches
to a fuel optimal formation keeping algorithm. This algorithm is based on the solution of

8
FUEL ERROR−BOX TRADE−OFF FOR VARIOUS D/DRAG VALUES
FUEL OPTIMAL FORMATIONKEEPING FOR 10 ORBITS 1
15 0.05e
−7

x0 : 5m y0 : 3m 2
ddrag = 1e−8 m/s 0.95 0.15e−7
−7
0.2e
−7
0.45e
10 0.9
0.8e−7

Normalized Fuel Usage For Each d/drag value


−7
1.35e
0.85 3.2e−7
−7
4.02e
5
0.8
RADIAL

0.75
0

0.7

−5 0.65

0.6
−10
0.55

−15 0.5
0 1 2 3 4 5 6 7 8 9 10
−20 −15 −10 −5 0 5 10 15 20
INTRACK 3 unit error box = +/− 20m intrack +/−4m radial

Fig. 10: Fuel Optimal Formation Keeping Fig. 11: Fuel Usage and Error Box Trade-
Under Disturbances off Under Differential Disturbances

an extension of the LP formation planning problem with the addition of differential dis-
turbances. Note that the very accurate CDGPS measurements could be used to produce a
parametric estimation of the constant and periodic components of the disturbance environ-
ment. Fig. 10 shows the results of a fuel-optimal formation keeping maneuver for 10 orbits
under a constant differential disturbance (drag) of 10−8 m/s2 .
Given a disturbance model (e.g. differential drag, differential J2 ), the fuel-optimal thrust-
ing sequence of a vehicle can be calculated to keep the vehicle within a desired error box
around a reference point or another vehicle. This is a complicated process for differential
J2 disturbances which are a function of the vehicle separations. However, if it is initially
assumed that the low-level control keeps the vehicles within the prescribed error boxes,
then the size of these boxes can be used to bound the maximum separation of the vehicles
and thus also the disturbances of the fleet. The low-level control design process can then
be iterated to develop new thrusting sequences that account for the updated differential
disturbance models.
As might be expected, the fuel cost for a zero tolerance error box is directly equal to
counter-acting the differential disturbance at all times. However, an interesting problem is
the trade-off analysis of fuel savings associated with larger error boxes. For this case we
can use the LP method to generate the data required to map the error box size against
fuel usage for a set of typical differential accelerations observed on on-orbit. Such a map is
shown in Fig. 11 for differential drag ranging from 4.02×10−7 m/s2 to 0.05×10−7 m/s2 . For
example, for a constant differential acceleration of 1.5×10−7 m/s2 , we could obtain almost a
30% fuel savings by changing from a 1 unit size error box to a 3 unit size error box (which
is ±20m in the intrack and ±4m in the radial direction). With analysis of this type, the
methods developed for formation keeping with disturbances can also be used to develop
insights on the overall mission design.
Both of the low-level schemes directly use the operation mode information from the
coordinator for selection of error box sizes. Table 1 shows the expected operation modes
for Orion with the corresponding separations and relative error boxes.
In addition, a phase-plane based approach is currently being extended to control in-

9
Experimental In-track Relative Radial
Modes Separation (m) Tolerance (m)
Coarse Parking 300 10-20
Fine Parking 100 ≤ 10
Precision Formation Flying 100 ≤2

Table 1: Separations and Tolerances for Orion Experiment Modes

dividual vehicles to desired formation locations. The approach divides the vehicle motion
into two parts that involve cyclic and secular motion. Simple control strategies can then
be developed to utilize differential energy correction and mean motion sizing to perform
formation keeping with a desired error box. The approach is similar in concept to the
eccentricity minimizing control in Ref. [29].
For both high-level coordination and low-level satellite algorithms, on-orbit estimation
plays a crucial role in the flight mode selection, system and environment parametric model-
ing, and the control input generation. The next section describes the measurement models
(absolute, relative and attitude) of the Carrier-phase Differential GPS (CDGPS) unit and
the estimation schemes for high precision formation flying.

Carrier-Phase Differential GPS

The sensing component of the control architecture is performed by the Carrier-phase Differ-
ential GPS sensing for precise absolute and relative navigation for formation flying [23, 24,
25]. The complete state of the formation of user satellites can be determined using GPS.
This includes the absolute position and velocity of each vehicle (generally determined in
Earth Centered Inertial frame), the relative position and velocity (the user vehicles with
respect to each other), and the attitude of the user vehicles that have sufficient antenna
and RF processing capability. This three-in-one capability of GPS sensing makes it very
attractive option, often able to replace three separate sensor systems, at a fraction of the
cost and weight.
The estimator produces three solutions: the attitude, the relative state between users,
and the absolute state of the (current) vehicle. The absolute state is determined by stan-
dard GPS pseudo-ranging, which measures the transmission time for the RF signals from
at least 4 NAVSTAR satellites. Using raw GPS data sent by direct crosslink from the other
satellites in the formation, and its own measurements, the relative state is determined using
carrier-phase differential GPS, where the phase within the 19.2cm carrier wave is measured
by two antennas, subtracted, and from that the relative distance can be measured. Atti-
tude is also determined by carrier-phase differential GPS, but uses multiple antennas at
fixed locations on a single vehicle frame. The method for determining relative position is
inherently about three orders of magnitude more accurate than the method for determining
absolute position. It is possible to use the known orbital dynamics in combination with the
much higher precision of the relative solution to improve the absolute position estimate.
Improved absolute position knowledge, especially in the radial direction, leads directly to
improved control performance.

As shown in Fig. 3, these solutions are transferred to the high-level coordinator. The
coordinator uses this information for both the high-level constellation management as well

10
−3 Differential Velocity
Differential Position x 10
0.1 1

0.5

radial (m/s)
0.05
radial (m)

0 0

−0.05 −0.5

−0.1 −1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
−3
x 10
0.1 2

in−track (m/s)
0.05 1
in−track (m)

0 0

−0.05 −1

−0.1 −2
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
−3
x 10
0.1 2

cross−track (m/s)
cross−track (m)

0.05 1

0 0

−0.05 −1

−0.1 −2
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Time (orbits) Time (orbits)

Fig. 12: Position Error for Kalman Filter Fig. 13: Velocity Error for Kalman Filter

as forwarding it down to the low-level satellite control. The next section discusses the
relative navigation estimator and the accuracy levels predicted.

Estimation

Note that both a Kalman filter and Weighted Least Squares approach are used for the
relative state estimation. For the Kalman filter, the model for the relative motion of the
vehicles in inertial space is used to propagate the current state and covariance matrix of
the vehicle to the next measurement time 1 Hz rate. This propagated state is then updated
with the measurements, where the model and measurements are combined using an optimal
weighting matrix. For this particular problem, the state was propagated with a non-linear
model, and the covariance matrix propagated with a linearized model. The Weighted least
squares is based entirely on the current measurements from a single time step. It weights
the measurements according to the expected noise levels on the signals.
Simulations were run to test the accuracy of these two estimators. The state of the GPS
NAVSTAR constellation was propagated over time, and measurements simulated from the
constellation. The simulation results shown assumed an orbit with 35 degree inclination,
400km altitude, and eccentricity 10−3 . Figs. 12, 13 show the position and velocity error
for Kalman filter relative state estimation between 2 vehicles with only an in-track initial
separation of 1000 meters. The noise model for the measurements was based on previous
actual GPS receiver performance. The absolute position ranging noise was assumed 25m,
the differential code phase noise (where common error sources, like S/A, are eliminated)
was assumed 2m, carrier noise 2cm, and Doppler noise 5mm/sec. Fig. 14 compares the
errors of the weighted least squares and Kalman filter solutions.
Multiple simulations were run with random noise included, and the final results from
these are in Table 2. Note that these are the averages over all simulations of the mean and
standard deviation on the absolute and relative errors for the second half of the orbit shown.
Note the units for each quantity. These results show the expected 2-5cm levels of accuracy
that should be achievable on-orbit in the relative position estimates at a 1 Hz update rate.

11
Comparison of Kalman and WLS
1

0.8
Pos Error (m)

0.6

0.4

0.2

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

0.02

0.015
Vel Error (m/s)

0.01

0.005

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Time (orbits)

Fig. 14: Comparison of Kalman and Weighted Least Square Filter Performances

PASSIVE APERTURES WITH MEAN FORMATION EC-


CENTRICITY
Many earth imaging and remote sensing applications would benefit from long periods of
operation over particular regions of interest, as is typical of Molniya-type orbits. In addi-
tion, highly eccentric orbits can be used to increase the percentage of useful observation
time by decreasing the occlusion time of the Earth [26]. Thus there is significant interest
in extending the design of “passive apertures” to the case of eccentric orbits, thereby pro-
viding a natural extension of the classic closed-form solutions to Hill’s equations. Using
the linearized relative equations of motion with respect to any Keplerian orbit (see [28] for
details), a generalization of the closed-form solutions [30] can be made for constellations
with a non-zero mean formation eccentricity. Ref. [28] also provides the necessary condi-
tions for obtaining T -periodic solutions for passive formations in eccentric orbits, where T
corresponds to orbital period.
Figs. 15 and 16 show a typical periodic closed-form solution for a reference orbit of
a=46000km, e=0.67, i =62.8◦ . The in-plane and out-of-plane motion correspond to incre-
mental changes in eccentricity (δ e = 0.0001) and inclination (δ i = 0.005◦ ). An interesting
feature of these eccentric orbits is the “figure 8” shaped out-of-plane motion. For circular

12
Error Mean St. Dev.
Absolute Position (m) 25.31 148.83
Relative Radial Position (cm) 2.73 2.34
Relative In-track Position (cm) 1.31 1.69
Relative Cross-track Position (cm) 1.28 0.82
Relative Radial Velocity (mm/s) 0.095 0.066
Relative In-track Velocity (mm/s) 0.018 0.051
Relative Cross-track Velocity (mm/s) 0.014 0.027

Table 2: Error Statistics for Estimation using a Kalman Filter

Fig. 15: In-Plane Relative Motion For a Fig. 16: Out-of-Plane Relative Motion For
High Eccentricity Reference Orbit, Nonlin- a High Eccentricity Reference Orbit, Non-
ear Simulation linear Simulation

reference orbits, small inclination differences only result in a one-dimensional out-of-plane


motion. This feature of highly eccentric orbits could be used to provide higher u-v plane
coverage per orbit for aperture filling observations [3]. Current efforts are focused on mod-
ifying the control design tools to account for mean formation eccentricities.

CONCLUSION
Formation flying is a key technology for deep space and orbital applications that involve
multi-spacecraft operations such as remote sensing and imaging applications. However,
under the overall formation flying problem, there are system-level challenges for such dis-
tributed systems due to limited resources and hardware constraints of space systems. This
paper describes ongoing research at Stanford University to address these systems-level is-
sues for two future on-orbit demonstrations of formation flying: Orion and TechSat21. A
particular emphasis of this work is on precise relative sensing and formation flying control
of multiple spacecraft via carrier-phase differential GPS.
The paper discusses a multi-layer autonomous formation flying control architecture.
This GNC tool combines low-level satellite control algorithms (formation keeping and rela-

13
tive error correction) with high-level fuel/time optimal formation coordination and planning
for multiple vehicles. The paper also discusses algorithms that can be used to perform the
high-level coordination, trajectory design, low-level satellite control, and CDGPS estimation
(absolute, relative and attitude). The basic features of this autonomous control architecture
was implemented on an independent nonlinear orbital simulation of 3 Orion vehicles with
realistic disturbances and a differential GPS measurement model.
Finally, we demonstrate that it is possible to obtain closed-form solutions for constella-
tions with mean formation eccentricity, which generalizes previous work on circular reference
orbits.

ACKNOWLEDGMENTS
The authors would like to thank Dr. Marc Jacobs (AFOSR), Rich Burns (AFRL), John
Bristow (NASA GSFC), and Dr. Frank Bauer (NASA GSFC) for their guidance and support
on the TechSat21 and Orion flight programs at Stanford University.

References
[1] D. Folta, L. Newman, T. Gardner, “Foundations of Formation Flying for Mission to Planet
Earth and New Millennium,” AIAA/AAS Astrodynamics Specialists Conference, July 1996.
[2] M. Colavita, C. Chu, E. Mettler, M. Milman, D. Royer, S. Shakian, and J. West, “Multiple
spacecraft interferometer constellation (MUSIC),” Tech. Rep. JPL D-13369, JPL - Advanced
Concepts Program, February 1996.
[3] C. A. Beichman, M. Shao, M. Colavita, and L. A. Lemmerman, ”Terrestrial planet finder
homepage,” http://eis.jpl.nasa.gov/origins/missions/terrplfndr.html.
[4] T. Alfery, “Request for proposal for deep space 3 (DS-3)spacecraft system industry partner,”
Tech. Rep. RFP No. N01-4-9048-213, Jet Propulsion Laboratory, November 1998.
[5] H. Hirosawa, H. Hirabayashi, T. Orii and E. Nakagawa, “Design of the Space-VLBI Satellite
MUSES-B,” 19th Int. Symposium on Space Technology and Science, Yokohama, May 1994.
[6] Timothy H. Dixon, “SAR Interferometry and Surface Detection,” Workshop Report, Boulder-
Colorado, Feb. 1994.
[7] R. S. Wimokur, “Operational Use of Civil Space-Based Synthetic Aperture Radar,” NOAA
JPL Publication 96-16, Aug 1996.
[8] E.M.C. Kong and D.W. Miller, ”Optimization of Separated Spacecraft Interferometer Trajec-
tories in the Absence of a Gravity-Well,” SPIE-3350-13, in Astronomical Interferometry Con-
ference Proceedings, 1998.
[9] R.J. Sedwick, T.L. Hacker and D.W. Miller, “Optimum Aperture Placement for a Space-
Based Radar System Using Separated Spacecraft Interferometry,” in Proc. of the AIAA GNC,
Aug. 1999.
[10] F. Bauer, J. Bristow, D. Folta, K. Hartman, D. Quinn, and J. P. How, “Satellite formation
flying using an innovative autonomous control system (AUTOCON) environment,” in Proc. of
the AIAA GNC, Aug 1997.
[11] S. Vadali, X Young, H. Schaub and K. T. Alfriend, “Fuel Optimal Control for Formation Flying
of Satellites,” presented at AIAA Guidance,Navigation and Control Conf., Aug. 1999.
[12] B. Morton, N. Weininger and J. Tierno, “Optimum Aperture Placement for a Space-Based
Radar System Using Separated Spacecraft Interferometry,” in Proc. of the AIAA GNC,
Aug. 1999.
[13] J. How, R. Twiggs, D. Weidow, K. Hartman, and F. Bauer, “Orion: A low-cost demonstration
of formation flying in space using GPS,” in AIAA Astrodynamics Specialists Conf., Aug 1998.

14
[14] D.C. Folta and D.A. Quinn, “A Universal 3-D Method for Controlling the Relative Motion of
Multiple Spacecraft in any Orbit,” in Proc. AIAA/AAS Astrodynamics Specialists Conf., Aug
1998.
[15] V. Kapila, A. G. Sparks, B. James and Q. Yan, “Spacecraft Formation Flying: Dynamics and
Control,” in Proc. of the ACC, June 1999.
[16] M. Meshabi and F. Y. Hadaegh, “Formation Flying Control of Multiple Spacecraft: Graph
Theoretic Properties and Switching Schemes,” in Proc. of the AIAA GNC, Aug. 1999.
[17] G.Q. Xing, S.A Parvez and D.C.Folta, “Implementation of Autonomous GPS Guidance and
Control for Spacecraft Formation,” in Proc. of the ACC, June 1999.
[18] AFRL Space Vehicles Directorate, http://www.vs.afrl.af.mil/factsheets/TechSat21.html.
[19] P.K.C. Wang and F.Y. Hadaegh, “Coordination and Control of Multiple Microspacecraft Mov-
ing in Formation,” Journal of Astronautical Sciences, Vol 44, No. 3, 1996, pp. 315-355.
[20] R.W. Beard, T.W. McLain and F.Y.Hadaegh, “Fuel Equalized Retargeting for Separated Space-
craft Interferometry,” in Proc. of ACC, June 1998.
[21] A. Robertson, G. Inalhan, and J. P. How, “Formation Control Strategies for a Separated Space-
craft Interferometer,” in Proc. of 1999 ACC, (San Diego, CA), June 1999.
[22] A. Robertson, G. Inalhan, and J. P. How, “Spacecraft Formation Flying Control Design for the
Orion Mission,” in Proceedings of AIAA/GNC, August 1999.
[23] A. Robertson, T. Corazzini, and J. P. How, “Formation sensing and control technologies for a
separated spacecraft interferometer,” in Proceedings of ACC, June 1998.
[24] C. Adams, A. Robertson, K. Zimmerman, and J. P. How, “Technologies for Spacecraft Forma-
tion Flying,” in Proceedings of the ION GPS-96 Conf., Sep 1996.
[25] T. Corazzini, A. Robertson, J. C. Adams, A. Hassibi, and J. P. How, “GPS Sensing for Space-
craft Formation Flying,” in Proceedings of the ION GPS-97, Sep 1997.
[26] Chandra X-Ray Observatory Mission Facts, http://chandra.harvard.edu/
[27] J. Hauser, D. G. Meyer, “Trajectory Morphing for Nonlinear Systems,” in Proceedings of the
ACC, June 1998.
[28] G. Inalhan, J. P. How, “Relative Dynamics & Control of Spacecraft Formations in Eccentric
Orbits,” submitted to The Journal of Astronautical Sciences, January 2000.
[29] C. L. Leonard, W. M. Hollister, E. V. Bergmann, “Orbital Formationkeeping with Differential
Drag,” Journal of Guidance, Control and Dynamics, Vol.12 No.1, Jan.-Feb. 1989.
[30] T. E. Carter, “New Form for the Optimal Rendezvous Equations Near a Keplerian Orbit,”
Journal of Guidance, Control and Dynamics, Vol.13 No.1, Jan.-Feb. 1990.
[31] A.I.Solutions, ”FreeFlyer User’s Guide”,Version 4.0, March 1999.

15

You might also like