ArticlePDF Available

Optimal Fuel-Balanced Impulsive Formationkeeping for Perturbed Spacecraft Orbits

Authors:

Abstract

This paper develops an impulsive spacecraft formation-flying control algorithm using relative-orbital-element corrections. This formalism introduces an inherent freedom that is used for deriving an optimal formation keeping law, balancing the fuel consumption among the spacecraft based on the impulsive Gauss variational equations. The main idea is that formulating the problem of formationkeeping in terms of relative-orbital-element corrections leaves the final values of the orbital elements unconstrained, thus allowing the spacecraft to create natural energy-balanced formation. The freedom rendered by this modeling is used to find optimal impulsive maneuvers, minimizing the squared l(2)-norm of the velocity-correction vector, which can be used for formation initialization and control. The optimization is solved using the least-squares method. The optimal formationkeeping method is designed to accommodate the effects of oblateness and drag. Based on graph theory, it is shown that the spacecraft will naturally form a stable energy-balanced formation and that the optimal formationkeeping strategy is invariant to the spanning tree. The algorithm is illustrated by simulating the motion of a formation of spacecraft possessing different ballistic coefficients subject to oblateness and drag.
Optimal Fuel-Balanced Impulsive Formationkeeping
for Perturbed Spacecraft Orbits
Igor Beigelmanand Pini Gurl
TechnionIsrael Institute of Technology, 32000 Haifa, Israel
DOI: 10.2514/1.34266
This paper develops an impulsive spacecraft formation-ying control algorithm using relative-orbital-element
corrections. This formalism introduces an inherent freedom that is used for deriving an optimal formationkeeping
law, balancing the fuel consumption among the spacecraft based on the impulsive Gauss variational equations. The
main idea is that formulating the problem of formationkeeping in terms of relative-orbital-element corrections leaves
the nal values of the orbital elements unconstrained, thus allowing the spacecraft to create a natural energy-
balanced formation. The freedom rendered by this modeling is used to nd optimal impulsive maneuvers,
minimizing the squared l2-norm of the velocity-correction vector, which can be used for formation initialization and
control. The optimization is solved using the least-squares method. The optimal formationkeeping method is
designed to accommodate the effects of oblateness and drag. Based on graph theory, it is shown that the spacecraft
will naturally form a stable energy-balanced formation and that the optimal formationkeeping strategy is invariant
to the spanning tree. The algorithm is illustrated by simulating the motion of a formation of spacecraft possessing
different ballistic coefcients subject to oblateness and drag.
Nomenclature
A= system matrix
a= semimajor axis
CD= drag coefcient
e= eccentricity
f= true anomaly
G= graph
g0= gravitational acceleration at sea level
H= cost functional
h= orbital angular momentum
I= inclination
In= modied Bessel function of the rst kind
Isp = specic impulse
KD= ballistic coefcient
M= mean anomaly
m= mass
n= mean motion
p= semilatus rectum
r= orbit radius
Si= spacecraft i
s= reference area
= classical orbital-element vector
= inverse of the atmospheric scale factor
v= velocity-correction vector
:= variation of :
= Lagrange multiplier vector
= gravitational constant
= right ascension of the ascending node
!= argument of perigee
kk
m=m-norm of a vector
I. Introduction
CONTROL of relative spacecraft motion, known as
formationkeeping, is an enabling technology for spacecraft
formation-ying missions. Modeling relative-motion dynamics, an
essential infrastructure for formationkeeping, has seen signicant
progress in recent years, since Clohessy and Wiltshire [1] published
their linearized relative-motion approximation in the early 1960s.
For example, Karlgaard and Lutze [2] derived an approximate
solution to second-order relative-motion equations for spacecraft in
near-circular Keplerian orbits. Gurl and Kasdin [3] presented a
methodology for obtaining high-order approximations of the relative
motion between spacecraft by using the Cartesian conguration
space in conjunction with the classical orbital elements. Chichka [4]
used the linear ClohessyWiltshire equations to nd natural
formations of satellites with a constant apparent distribution, an
important application of the relative-motion equations for remote
sensing.
However, natural formations may not survive, due to orbital
perturbations. The need for formationkeeping, therefore, stems from
the fact that space is a perturbed environment, thus inducing
separation of the formation-ying spacecraft orbital planes. In low
Earth orbits, the dominant perturbations are the Earth oblateness and
drag, and at higher altitudes, the spacecraft are perturbed by the
attraction forces of the sun and the moon.
The growing number of autonomous spacecraft missions that are
to perform complicated missions in close proximity requires
accurate, robust, and fuel-efcient formationkeeping algorithms.
Thus, attention should be given to developing a new rigorous control
approach, taking into account the effect of all dominant orbital
perturbations.
One such perturbation is Earths oblateness. Because J2
perturbations affect the orbital elements of each spacecraft in the
formation differently, the formation will tend to separate. An
analytical method developed by Alfriend et al. [5] can be used to
evaluate the differential forces on the formation-ying spacecraft.
Moreover, Schaub and Alfriend [6] also found that J2-invariant
relative orbits can be designed analytically; spacecraft placed on the
SchaubAlfriend orbits will be inuenced by the same forces and
will therefore have the same drift, thus yielding bounded relative
motion in the conguration space.
Formationkeeping can be implemented by using either impulsive
control, relying on chemical thrusters, or continuous control, using
low-thrust electric propulsion. Impulsive formationkeeping usually
assumes that the orbital elements of some reference orbit are known
Presented as Paper 6544 at the AIAA Guidance, Navigation and Control
Conference and Exhibit, Hilton Head, SC, 2023 August 2007; received 26
August 2007; revision received 8 January 2008; accepted for publication 8
January 2008. Copyright © 2008 by The Authors. Published by the American
Institute of Aeronautics and Astronautics, Inc., with permission. Copies of
this paper may be made for personal or internal use, on condition that the
copier pay the $10.00 per-copy fee to the Copyright Clearance Center, Inc.,
222 Rosewood Drive, Danvers, MA 01923; include the code 0731-5090/08
$10.00 in correspondence with the CCC.
Graduate Student, Faculty of Aerospace Engineering; beigel@tx.
technion.ac.il.
Senior Lecturer, Faculty of Aerospace Engineering; pgurl@technion.
ac.il. Associate Fellow AIAA.
JOURNAL OF GUIDANCE,CONTROL,AND DYNAMICS
Vol. 31, No. 5, SeptemberOctober 2008
1266
Downloaded by TECHNION - ISRAEL INST OF TECH on December 19, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.34266
and attempts to generate control commands that will match the
instantaneous orbital elements to some desired values [7], even under
orbital perturbations [8]. However, in many cases, such as in situ eld
mapping and astronomical observations, setting an a priori reference
orbit is not a mission requirement (although orbit knowledge is
required). More important, on some occasions, specifying the
reference orbit a priori can result in a greater fuel consumption.
At this point, one may ask what can be done to the improve the
existing [710] impulsive-formationkeeping schemes. The answer is
that the overall amount of propellent used for formationkeeping can
be optimally distributed among the spacecraft if the reference orbit
for the formation-ying spacecraft is not specied. In other words, if
one used the relative-orbital-element corrections to dene the
formationkeeping strategy, the formation could be designed so that
the relative geometry is determined without imposing an absolute
conguration on the formation. This approach is conceptually
different from existing distributed spacecraft system optimization
methods [11].
Relative-orbital-element corrections are the differences between
the successive corrections of orbital elements of any two formation-
ying spacecraft. In essence, these elements differ from the orbital-
element differences proposed by Schaub [12], because they do not
depend on a particular reference orbit. The proposed modeling
approach remains valid in the presence of orbital perturbations.
In this work, we develop a generic method for the impulsive
control of formation-ying spacecraft using the Gauss variational
equations (GVEs) [13] as the dynamic model and the classical
relative-orbital-element corrections as the variables representing the
formation structure. As opposed to other works [1416], we shall not
attempt to minimize the overall amount of fuel consumed by the
formation-ying spacecraft, but rather to optimally balance a given
amount of fuel among the spacecraft. This prevents an undesirable
situation in which some of the spacecraft maneuver much more than
others, thus considerably increasing the a priori propellent margin of
the entire formation.
This approach relies on the key observation that formulating the
problem of formationkeeping in terms of relative-orbital-element
corrections leaves the nal values of the corrected elements
unconstrained. This freedom can be used to design optimal impulsive
maneuvers for formation initialization, reconguration and
maintenance. We show that our impulsive control scheme is a
natural scheme in the sense that it creates a stable formation that
corresponds to an optimum balancing of the overall amount of
propellent. This approach is substantially different from other fuel-
balancing methods [17], wherein the control design exploits the
approximate dynamics given by the ClohessyWiltshire equations
and an approximate perturbation model.
In our analysis, we use elementary graph theory [18] to classify the
formation topologies and to guide the quest for optimal maneuvers.
Graph theory is a useful tool for analyzing multiple spacecraft
formations; thus far, however, it has been mostly used in the
continuous-formationkeeping setting [19]. We offer to apply a
graph-theory-inspired analysis for impulsive-formationkeeping
design.
II. Dynamic Model
In this section, we briey outline the underlying dynamic model
used for the design of optimal formationkeeping, including the
effects of J2and drag.
A. Coordinate Systems
We rst dene the following coordinate systems [20]:
1) Ris a rotating polar coordinate system, centered at the
spacecraft. The fundamental plane is the orbital plane. The unit
vector ^
ris directed from the spacecraft radially outward, ^
his normal
to the fundamental plane, and ^
completes the right-hand setup.
2) Tis a rotating tangential-normal frame, centered at the
spacecraft. The fundamental plane is the orbital plane. The unit
vector ^
tlies along the spacecraft velocity vector, ^
hcoincides with the
instantaneous angular momentum vector, and ^
ncompletes the right-
hand setup.
3) Pis a perifocal coordinate system, centered at the primary. The
fundamental plane is the orbital plane. The unit vector ^
xpis directed
from the primarys center to the periapsis, ^
zpis normal to the
fundamental plane, and ^
ypcompletes the right-hand setup.
4) Iis an inertial coordinate system, centered at the Earth. The
fundamental plane is the equator, ^
Xis directed from the Earths
center to the vernal equinox, ^
Zis normal to the fundamental plane,
and ^
Ycompletes the right-hand setup.
B. Gauss Variational Equations
The GVEs model the effect of a control and/or a disturbance
acceleration vector
uut;u
n;u
hT
on the osculating orbital elements time derivatives. This vector is
represented in the reference frame T, and so utand unare the input
components in the plane of the osculating orbit along the velocity
vector and perpendicular to it, respectively, and uhlies along the
instantaneous angular momentum vector, normal to the orbital plane.
In the case of impulsive maneuvers, we can write
_
t; uttVt;u
ntVn;
uhtVh
(1)
where is an orbital element; is an orbital-element correction;
Vt,Vn, and Vhare the components of the velocity-correction
vector in frame T; and the impulsive time interval t!0. The
impulsive form of the GVEs denes algebraic relationships between
the orbital-element corrections and the components of the velocity-
correction vector [20]:
a2a2v
Vt(2a)
e1
v2ecos fVtr
asin fVn(2b)
Ircosf!
hVh(2c)
 rsinf!
hsin IVh(2d)
!1
ev 2 sin fVt2er
acos fVn
rsinf!cos I
hsin IVh(2e)
M b
eav 21e2r
psin fVtr
acos fVn(2f)
where ais the semimajor axis, eis the eccentricity, Iis the
inclination, is the right ascension of the ascending node, !is the
argument of perigee, Mis the mean anomaly, the parameter fis the
true anomaly, ris the scalar orbit radius, and pis the semilatus
rectum. The mean motion nsatises n
=a3
p, where is the
gravitational parameter. The semiminor axis bsatises
ba
1e2
p.
Consequently, in matrix form, GVEs (2) for some spacecraft Si
can be written as
BEIGELMAN AND GURFIL 1267
Downloaded by TECHNION - ISRAEL INST OF TECH on December 19, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.34266
iGivi(3)
where
i
a
e
I

!
M
2
6
6
6
6
6
6
4
3
7
7
7
7
7
7
5i
;Gi
cT
a
cT
e
cT
I
cT
cT
!
cT
M
2
6
6
6
6
6
6
4
3
7
7
7
7
7
7
5i
;vi
Vt
Vn
Vh
2
43
5i
(4)
where cTdenotes a coefcients vector of the controlled orbital
elements emanating from the GVEs, as seen in Eqs. (2), and vis the
concomitant impulsive velocity-correction vector.
C. Mean Orbital Elements Subject to a J2Perturbation
In the absence of perturbations, the six orbital elements remain
constant. Because of the inuence of J2, some orbital elements will
oscillate about a nominal value and others will exhibit secular drifts.
The averaged values of the orbital elements are called the mean
elements. If all the spacecraft in the formation are of equal type and
build (i.e., have the same ballistic coefcient), then the differential J2
acceleration is the dominant perturbation experienced by the
spacecraft. In this case, the differential drag effect on the relative
motion is negligible over a time period of several orbits.
The mean orbital elements under the effect of J2are determined by
the differential equations [20]:
da
dt0(5)
de
dt0(6)
di
dt0(7)
d
dt3
2J2Re
p2
ncos I(8)
d!
dt3
4J2Re
p2
n5cos2I1(9)
dM
dtn3
4J2Re
p2
n
1e2
p3cos2I1(10)
where Reis the equatorial radius.
D. Mean Change of the Orbital Elements Because of Drag
If atmospheric rotation is neglected, then the inclination, the right
ascension of the ascending node, and the argument of perigee are not
affected by drag. The relationships for the averaged rates of the
semimajor axis and the eccentricity, developed for a small
eccentricity (expanding up to second-order in e), are given by [8]
da
dt2pa2nKDI02eI13e2
4I0I2expae
(11)
de
dtpKDan2I1eI0I2e2
45II1I3expae
(12)
where Ikare modied Bessel functions of the rst kind of order kand
argument ae,dened in the integral form,
Ikz 1
2Z2
0
ezcos coskd(13)
pis the atmospheric density at perigee, is the inverse of the
atmospheric scale height, and the ballistic coefcient is
KDsCD
2m
where CDis the drag coefcient, mis the spacecraft mass, and sis a
reference area.
E. Relative Position and Velocity in Inertial Coordinates
To model the relative motion of the spacecraft under the preceding
orbital perturbations, we will rst write the inertial position and
velocity of each spacecraft. This can be done by solving the
Keplerian two-body problem in inertial coordinates and then
extending the solution to include orbital perturbations using a
standard variation-of-parameters procedure. The inertial position r
and inertial velocity v_
rare given by [20]
ra1e2
1ecos f
cosf!cos  cos Isinf!sin 
cos Icos  sinf!cosf!sin 
sin Isinf!
2
43
5
(14)
v_
r
a1e2
rVx;V
y;V
zT(15)
where
Vxcos  sinf!sin  cos Icosf!
ecos  sin !sin  cos !cos I
Vycos  cos Icosf!sin  sinf!esin  sin !
cos  cos !cos I
Vzsin Icosf!ecos !
In the presence of orbital perturbations and velocity corrections,
the classical orbital elements vary with time. In this case, the inertial
position and velocity are determined as follows:
1) Integrate the orbital elements with respect to time under orbital
perturbations.
2) Calculate the inertial velocity until the moment of velocity
correction.
3) Compute the inertial position by quadrature: rR_
rdt.
4) Perform an impulsive correction v(the required velocity-
correction components are transformed from frame Tto I).
5) Determine the inertial velocity after the impulsive correction:
vvv.
6) Integrate the orbital elements (step 1).
The relative velocity and relative position vectors in frame I
between any two spacecraft Sjand Siare dened as follows:
_
ri;j Vxi;j;Vy
i;j;Vz
i;jT_
rj_
ri(16)
ri;j Xi;j;Y
i;j;Z
i;jTrjri(17)
III. Formation Topologies and Optimal
Formationkeeping
In this section, we consider the relative motion of a group of N
spacecraft Si(i1;...;N). We will develop an impulsive-
formationkeeping maneuver strategy for consuming minimum fuel.
1268 BEIGELMAN AND GURFIL
Downloaded by TECHNION - ISRAEL INST OF TECH on December 19, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.34266
The impulsive maneuver will be represented by a velocity-correction
vector performed by each spacecraft, and the formation topology will
be represented using elementary graph theory, briey described in
the following subsection.
A. Elementary Graph Theory
A graph is a pair [GV; E] of sets such that the elements of E
are two-element subsets of V. The elements of Vare the vertices or
nodes, and the elements of Eare the edges or arcs. The degree of a
vertex v,dGv, is the number of edges at v. A labeled graph is a
graph with each node labeled differently (but arbitrarily), so that all
nodes are considered to be distinct for the purpose of enumeration. A
path PV; Eon a labeled graph is a sequence
Vfv1;v
2;...;v
Ng
such that
Efv1v2;v2v3;...;vN1vNg
are the graph edges of the graph GV; E. A closed path
Pv1;v
2;...;v
N;v
1
with N3is called a graph cycle. The length of a cycle is its number
of edges; a cycle of length mis denoted by Cm. A directed graph, or
digraph, Gis an ordered pair [GV; A], where Vis a set of
vertices or nodes and Ais a set of ordered pairs of vertices (called
directed edges, arcs, or arrows). An edge ex; yis considered to
be directed from xto y(yis called the head and xis called the tail of
the edge); yis said to be a direct successor of x, and xis said to be a
direct predecessor of y. If a path leads from xto y, then yis said to be a
successor of x, and xis said to be a predecessor of y. A tree Tis a set of
straight-line segments connected at their ends, containing no closed
loops (cycles). In other words, it is a simple, undirected, connected,
acyclic graph. A tree with Nnodes has N1graph edges.
Conversely, a connected graph with Nnodes and N1edges is a
tree. A labeled tree is a tree in which each vertex is given a unique
label. The vertices of a labeled tree with nvertices are typically given
the labels 1;2;...;n. The leader graph is denoted by Lk; the leader
graph is a tree, which is a directed graph. A spanning tree of a
connected graph Gis either the maximal set of edges of Gthat
contains no cycles or the minimal set of edges that connect all
vertices.
B. Relative-Orbital-Element Corrections
To develop an impulsive control for a group of Nspacecraft, we
shall treat this group as a digraph GNand will term it the formation
graph. Each spacecraft Si(i1;...;N) constitutes a vertex in GN.A
relative-orbital-element correction of Siand Sj, denoted by i;j ,
constitutes an edge in GNand is dened by
i;j ji(18)
where the orbital-element corrections of Siand Sjin the presence of
an impulsive velocity correction are, respectively,
i
i
i(19)
j
j
j(20)
where is a value of the orbital element before an impulsive
correction, and is the value of an orbital element after the
impulsive correction. Therefore, a relative-orbital-element correc-
tion can be written as
i;j ji
j
j
i
i(21)
The relative-orbital-element corrections are indifferent to a particular
predened reference value chosen for an orbital element, ref (i.e.,
they are reference-orbit-independent); this can be seen by writing
i
iref ;j
jref (22)
and
i;j ji
j
i(23)
If the corrected orbital elements of Siare to be equal to the orbital
elements of Sj, then in the absence of perturbations, a formation will
be created; this case is referred to as perturbation-free orbital-element
matching. Semimajor axis (energy) matching in the case of Keplerian
motion is a typical example [10]:
a
ia
j)ai;j a
ia
j(24)
If perturbations are present, their effect will generally induce a
relative drift. The basic requirement from the formation control law is
to cancel the mean relative drift. When the relative deviation exceeds
a maximum allowed value, a velocity correction should be applied.
Each velocity correction sets the current value of the orbital-element
correction to a new value that cancels future drifts. In this case, the
corrected orbital elements of Siand Sjsatisfy the following
condition:
j
i
i;j (25)
and the relative element correction is
i;j
i;j
i;j (26)
where
i;j is the required relative orbital element of Siand Sj,dened
as
i;j ji
after the impulsive correction (this correction is determined based on
the effect of drag and J2, as will be shown in the sequel). Similarly,
i;j is the relative orbital element before the impulsive correction.
C. Tree Topology
Let us assume for the moment that there are no perturbations
present and that the formation is created by orbital-element
matching. The spacecraft form a leader graph (i.e., a tree), as dened
in Sec. III.A. Let us label the leader of this graph by S1, with SNbeing
the root. Per this denition, the spacecraft form a leader formation
LN, and the resulting formation topology assumes the following
form:
1;2
2;3
.
.
.
N1;N
2
6
6
6
4
3
7
7
7
5
110 0
011 0
   
00 1 1
2
6
6
43
7
7
5
1
2
.
.
.
N
2
6
6
6
4
3
7
7
7
5
(27)
Using GVEs (2), we note that the orbital-element corrections can be
expressed as a function of the velocity correction in the following
manner:
icT
ivi;jcT
jvj(28)
where ciand cjare GVE coefcient vectors of Siand Sj,
respectively, as seen in Eqs. (2), and vi,vjare the concomitant
impulsive velocity-correction vectors. The relative-orbital-element
correction can be thus written as follows:
i;j jicT
jvjcT
ivi(29)
The state-space representation of this tree can be written as
Av b(30)
where
BEIGELMAN AND GURFIL 1269
Downloaded by TECHNION - ISRAEL INST OF TECH on December 19, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.34266
A
cT
1cT
20 0
0cT
2cT
3 0
   
00 cT
N1cT
N
2
6
6
43
7
7
5
(31)
is an N13Nmatrix, and
b
1;2
2;3
.
.
.
N1;N
2
6
6
6
4
3
7
7
7
5
(32)
is a N11vector.
D. Balancing the Fuel Consumption
The use of relative-orbital-element corrections as edges in the tree
topology has introduced an excess freedom into the formation graph;
this can be plainly seen by the fact that the number of unknowns in
Eq. (30) exceeds the number of constraints. This implies that a static
parameter optimization problem with equality constrains can be
solved; the only remaining issue is to choose the proper cost
functional.
Choosing a cost functional that minimizes the total vof the
formation seems to be the most straightforward selection [1416].
For a formation of spacecraft equipped with separate impulsive
thrusters for all three axes, this cost functional is [21]
v
v
1
X
N
i1
vi
1
(33)
For spacecraft equipped with gimbaled thrusters, the minimum-fuel
cost functional for the formation is
v
v
2
X
N
i1
vi
2
(34)
In both cases, the cost functional constitutes the overall vof the
formation. The rationale for choosing the overall vas the cost stems
from the fact that the vis proportional to the amount of propellant
m. This can be readily seen from the rocket equation:
mm01expv
Ispg0 (35)
where m0is the initial fuel mass, Isp is the specic impulse, and g0is
the gravitational acceleration at sea level. For a rst-order small m,
Eq. (35) can be written as
vmIspg0
m0Om2(36)
However, minimizing either Eq. (33) or Eq. (34) may lead to a
situation wherein some of the spacecraft consume much more
propellant than the others. This unbalance can cause some of the
spacecraft to run out of fuel faster than the others; if the formation is
to be designed with identical or nearly identical spacecraft, each
spacecraft must be designed for the worst-case fuel consumption,
thus resulting in a conservative fuel budget for the entire formation.
To alleviate this situation, we propose using a penalty not on m,
but rather on m2. Although not necessarily minimizing the overall
fuel consumption, this penalizes large individual fuel usage and
balances the fuel among all the formation members. In this case, the
spacecraft will tend to form a natural fuel-balanced formation: that is,
to converge onto an orbit that does not require excessive fuel
consumption from some of the formation members while preventing
other spacecraft from maneuvering.
Penalizing m2implies penalizing v2; again, this can be seen
from Eq. (35):
v2m2I2
spg2
0
m2
0Om3(37)
Thus, the problem of fuel-balanced optimal formationkeeping can be
cast as follows. Find an optimal impulsive maneuver v, satisfying
varg min
v
v
2
2arg min
vX
N
i1
vi
2
2
(38)
such that
Av b(39)
Augmenting the cost functional with the equality constrains using the
Lagrange multiplier vector
1;
2;...;
N1T
yields
Hkvk2
2TAvb(40)
The necessary and sufcient conditions for the existence of minima
are
@H
@v0(41)
@2H
@v2>0(42)
The Hessian appearing in Eq. (42) is
@2H
@v22IN(43)
where INis an NNidentity matrix. Thus, the solution of Eq. (41) is
always a minimum, meaning that the solution of the optimization
problem (38) and (39) yields an optimal impulsive velocity-
correction vector v. This observation holds regardless of the
particular spanning tree used to dene the formation, as shown in
Appendix A.
A straightforward approach for solving the optimization problem
formulated in Eqs. (38) and (39) is to use the least-squares method.
To that end, we rewrite expression (39) into
AvbAvbTAvAvTbAvTAvbT
bbTAvvTAT2bvTATbbT(44)
Taking the derivative with respect to vyields
2AATv2ATb0(45)
Therefore, the minimizing vector vis a solution of the equation
AATvATb(46)
If the rows of Aare linearly independent, then AATis invertible. In
that case, the optimal solution of the system of linear equations is
unique and is given by
vATAAT1bAb(47)
where Ais the pseudoinverse of A.
IV. Control Strategy
The rst goal of the control strategy is to initialize the spacecraft
formation and then to keep the spacecraft in the leader graph LNeven
in the presence of perturbations. In the absence of perturbations, the
necessary and sufcient condition guaranteeing a stable formation is
that the mean motions of Siand Sjform a 1:1 resonant motion.
Because the periods of Keplerian elliptic orbits are determined by the
1270 BEIGELMAN AND GURFIL
Downloaded by TECHNION - ISRAEL INST OF TECH on December 19, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.34266
orbital energy, this requirement can be transformed into an energy
matching condition [10] or, equivalently, semimajor-axis matching.
However, orbital perturbations induce temporal changes in the
orbital elements of each spacecraft. In general, the effects of
perturbations on each spacecraft are not the same. The Earth
oblateness effect is a function of the semimajor axis, eccentricity, and
inclination. Because of initialization errors, small initial differences
of the semimajor axis, eccentricity, and inclination lead to secular
changes in both the nodal rate _
and the mean latitude rate _
M_
!.
These differences lead to a buildup of out-of-plane and in-plane
angular separations. In addition, not all the spacecraft involved in the
formation are of equal type, mass, shape, and attitude prole. These
physical variations lead to drag differences that cause secular
changes in the semimajor axis and the eccentricity, which in turn lead
to secular changes in the nodal rate and the mean latitude rate due to
the Earth oblateness.
The formationkeeping control law is required to mitigate the
relative deviations of the mean drift of _
and _
M_
!. That could be
done by persistently matching semimajor axis, eccentricity, and
inclination.
A. Optimal Corrections of the Semimajor Axis, Eccentricity, and
Inclination
Two different scenarios should be considered. The rst is when
all the spacecraft in the formation experience the same drag; the
relative drift of the orbital elements is then zero. The second scenario
is when the effect of natural perturbations on each spacecraft is
different.
To preserve the formation, a periodic velocity correction should
be applied. As already discussed, the corrected elements a
i,e
i, and
I
i(i1;...;N) would theoretically remain unchanged in the
absence of drag perturbations, and no further correction of these
elements would be needed. However, when a drag difference exists,
the elements a
i,e
i, and I
i(i1;...;N) slowly deviate from the
initial values until an additional velocity correction must be applied.
The time interval between subsequent corrections is chosen
according to the allowed deviation. When the deviation exceeds
the allowed maximum, a new velocity correction is performed.
Each velocity correction sets the element to cancel future drifts due
to drag.
Choosing to control semimajor axis, eccentricity, and inclination
leads to a relationship in the form of Eq. (30). Let us dene the
matrices Aa,Ae, and AIrepresenting the mapping of the velocity
corrections to the relative-orbital-element corrections ai;j ,ei;j,
and Ii;j (i1;...;N) based on GVEs (2):
Aa
ca1ca20 01N01N
0ca2ca3 01N01N
.
.
...
..
.
...
..
.
.
0 caN1caN01N01N
2
6
6
6
4
3
7
7
7
5
(48)
where caiis given by [compare with Eqs. (2)]
cai2via2
i
(49)
Similarly,
Ae
ce1ce20 se1se20 01N
0ce2ce3 0se1se3 01N
.
.
...
...
...
...
..
.
..
.
.
2
6
43
7
5
(50)
where
cei2eicos fi
vi
;seirisin fi
viai
(51)
and
Ai
01N01NkI1kI20
01N01N0kI2kI3
.
.
...
...
...
..
.
.
2
6
43
7
5(52)
where
kIiricosfi!i
hi
(53)
Finally, in the absence of perturbations, Eq. (30) assumes the
following form:
Aa
Ae
AI
2
43
5
Vt1
Vt2
.
.
.
VtN
Vn1
Vn2
.
.
.
VnN
Vh1
Vh2
.
.
.
VhN
2
6
6
6
6
6
6
6
6
6
6
6
6
6
6
6
6
6
6
6
6
6
4
3
7
7
7
7
7
7
7
7
7
7
7
7
7
7
7
7
7
7
7
7
7
5
a1;2
a2;3
.
.
.
aN;N1
e1;2
e2;3
.
.
.
eN;N1
I1;2
I2;3
.
.
.
IN;N1
2
6
6
6
6
6
6
6
6
6
6
6
6
6
6
6
6
6
6
6
6
6
4
3
7
7
7
7
7
7
7
7
7
7
7
7
7
7
7
7
7
7
7
7
7
5
(54)
which complies with the general form of Eq. (39).
When perturbations are present, the vector bbecomes a
superposition of the precorrected relative orbital elements and the
nal relative orbital elements, as shown in Eq. (26). Thus, the
formationkeeping topology developed here is generic; the only
difference between the perturbed and unperturbed cases is the entries
of b.
B. Closed-Form Expressions for Optimal Impulsive
Formationkeeping
Consider the following scenario: A formation graph G3is located
in nearby orbits with different initial semimajor-axis values. In the
absence of perturbations, this graph will form a stable formation if the
energies of S1,S2, and S3are equal or, equivalently, if the semimajor
axes of their orbits are identical. We will show that the preceding
optimal formationkeeping scheme can yield analytic closed-form
expressions for the optimal impulsive maneuvers required to match
the semimajor axes using the least-squares method.
The initial values of semimajor axes and velocities of G3are a1,
a2, and a3and v1,v2, and v3, respectively, and the relative semimajor
axes corrections are
a2;1a2a1(55)
a3;2a3a2(56)
The matrix Aa, which is given by Eqs. (48) and (49), assumes the
following form:
Aa2v1a2
1
2v2a2
2
0
02v2a2
2
2v3a2
3
2
43
5(57)
The optimal impulsive maneuver is calculated using the least-
squares method:
vAT
aAaAT
a1b;ba2;1;a3;2T(58)
The optimal velocity corrections required to match the semimajor
axes of the spacecraft are given by
BEIGELMAN AND GURFIL 1271
Downloaded by TECHNION - ISRAEL INST OF TECH on December 19, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.34266
V?
t11
2
v1a2
1v2
2a4
2a3;2v2
2a4
2a2;1v2
3a4
3a2;1
v2
1a4
1v2
3a4
3v2
1a4
1v2
2a4
2v2
3a4
3v2
2a4
2
(59)
V?
t21
2
v2a2
2v2
3a4
3a2;1v2
1a4
1a3;2
v2
1a4
1v2
3a4
3v2
1a4
1v2
2a4
2v2
3a4
3v2
2a4
2
(60)
V?
t31
2
v3a2
3v2
1a4
1a3;2v2
2a4
2a3;2v2
2a4
2a2;1
v2
1a4
1v2
3a4
3v2
1a4
1v2
2a4
2v2
3a4
3v2
2a4
2
(61)
and the resulting semimajor-axis corrections are
a?
i2a2
ivi
V?
ti(62)
yielding
a?
1
v2
1a4
1v2
2a4
2a3;2v2
2a4
2a2;1v2
3a4
3a2;1
v2
1a4
1v2
3a4
3v2
1a4
1v2
2a4
2v2
3a4
3v2
2a4
2
(63)
a?
2
a4
2v2
2v2
3a4
3a2;1v2
1a4
1a3;2
v2
1a4
1v2
3a4
3v2
1a4
1v2
2a4
2v2
3a4
3v2
2a4
2
(64)
a?
3
a4
3v2
3v2
1a4
1a3;2v2
2a4
2a3;2v2
2a4
2a2;1
v2
1a4
1v2
3a4
3v2
1a4
1v2
2a4
2v2
3a4
3v2
2a4
2
(65)
All the spacecraft in G3will match their semimajor axes to some
value a?. The value of a?corresponds to the balanced velocity
correction required to form a formation in the absence of
perturbations and is given by
a?aia?
i(66)
It can be shown that a?is the centroid of the initial values of the
semimajor axis:
a?1
NX
N
i0
ai(67)
This value of a?yields the globally optimal total velocity correction
in the sense of minimizing kvk2
2for all possible values of a
reference semimajor axis (this result is valid only for controlling each
orbital element separately).
V. Optimal Formationkeeping Under Perturbations:
Identical Ballistic Coefcients
In this section, we will implement the new control strategy on a
formation of J2-perturbed spacecraft having identical ballistic
coefcients. We will extend the method to the general case in the next
section. To illustrate the main idea, consider a formation of four
spacecraft ying in a low Earth orbit. All spacecraft have the same
ballistic coefcient and a small initial difference in the values of the
semimajor axis and eccentricity. In this case, as was described in
Sec. IV, to form a stable formation, one can match semimajor axis
and eccentricity, then the relative deviations of the mean drifts _
and
_
M_
!would be zero. The initial orbital elements and ballistic
coefcients are summarized in Table 1.
The initial differences in semimajor axis and eccentricity lead to a
buildup of angular separation between the orbital planes of the
spacecraft, i;j, and of the in-plane angular separation M!i;j ,
which increases the distance between spacecraft, di;j. We shall now
demonstrate the implementation of the proposed formationkeeping
method (matching semimajor axis, eccentricity, and inclination).
Figures 1 and 2 depict the time history of the mean relative orbital
elements. The impulsive maneuver is applied at f90 deg after
1.25 orbits. It can be seen that relative drift of the mean relative
orbital elements is stopped. Although the values of semimajor axis
and eccentricity decrease due to drag, the formation remains intact,
and the distances between spacecraft are bounded. Figure 3 shows
the time history of the relative velocity components Vxi;j,Vyi;j, and
Vzi;j in inertial coordinates [Eq. (16)]. The discontinuity in the
relative velocity components is a result of the impulsive velocity
correction. Following the impulsive maneuver, the position
components exhibit a bounded periodic motion. This can be seen
in Fig. 4, showing the three-dimensional relative orbit in inertial
coordinates and the projections thereof on the XY,XZ, and YZ
planes.
VI. Optimal Formationkeeping of Spacecraft with
Different Ballistic Coefcients
In this section, we study the motion of a formation of spacecraft
having different ballistic coefcients, and we develop an optimal
formationkeeping algorithm based on the least-squares formalism
developed in Sec. III. Different ballistic coefcients cause different
secular changes of semimajor axis and eccentricity for each
spacecraft, which in turn lead to different secular changes of the
nodal rate and the mean latitude rate, due to the Earth oblateness.
Periodic velocity corrections must then be applied, because the
corrected orbital elements will continue to deviate from the reference
values.
To that end, we will implement the optimal impulsive scheme
developed herein while relying on an auxiliary dynamic equations
developed by Mishne [8]. The auxiliary relations will be used to
calculate the required corrections of the semimajor axis and the
eccentricity that will cancel relative drifts in the nodal angle and the
mean latitude angle until the next velocity correction. We shall also
suggest a new method for mitigating the cross-coupling of the
orbital-element corrections (semimajor axis and eccentricity are
corrected by applying tangential and normal maneuvers; this
correction affects the mean latitude rate as well).
Before commencing with our development of the formationkeep-
ing strategy, we will highlight some important differences between
the current work and the method of Mishne [8]:
1) We propose controlling nodal rate and mean latitude rate
indirectly, by selecting suitable values of semimajor axis and
eccentricity. These values are chosen to cancel future drifts until the
next velocity correction. It is possible to choose which two of the
three elements to control because there are only two constraints:
namely, equalizing the out-of-plane and in-plane drifts.
2) Fuel optimization is done according to the method described in
Sec. III. Controlling semimajor axis and eccentricity leads to in-plane
velocity corrections only. By nding a suitable timing for these
corrections, it is possible to minimize the effect of changes in !and
Mdue to the impulsive corrections in the tangential and normal
directions.
3) We offer an optimal solution by not imposing a global reference
orbit a priori.
Table 1 Initial orbital elements and ballistic coefcients of a formation
in a low Earth orbit.
Sia,km eI, rad , rad !, rad f0, rad KD
S16928.2 0.0012 0.01 0 0 0 0.0355
S26928.3 0.0013 0.01 0 0 0 0.0355
S36928.5 0.0014 0.01 0 0 0 0.0355
S46928.8 0.0015 0.01 0 0 0 0.0355
1272 BEIGELMAN AND GURFIL
Downloaded by TECHNION - ISRAEL INST OF TECH on December 19, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.34266
0123456
−0.4
−0.3
−0.2
−0.1
0
0.1
ai,j [km]
Time [orbits]
0123456
−15
−10
−5
0
5x 10−5
ei,j
Time [orbits]
0123456
−0.03
−0.02
−0.01
0
0.01
fi,j [rad]
Time [orbits]
f2,1
f3,2
f4,3
e2,1
e3,2
e4,3
a2,1
a3,2
a4,3
Fig. 1 Mean values of the relative semimajor axis and eccentricity and the true anomaly, with an optimal velocity correction applied.
0123456
−2
0
2
4
6x 10−4
(ω+M)i,j [rad]
Time [orbits]
0123456
−2
−1.5
−1
−0.5
0x 10−6
i,j [rad]
Time [orbits]
2,1
3,2
4,3
(ω+M)2,1
(ω+M)3,2
(ω+M)4,3
Fig. 2 Relative drift in !Mand is arrested after a velocity correction.
BEIGELMAN AND GURFIL 1273
Downloaded by TECHNION - ISRAEL INST OF TECH on December 19, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.34266
0123456
−4
−2
0
2
4x 10−3
Vxi,j [km/sec]
Time [orbits]
Vx2,1
Vx3,2
Vx4,3
0123456
−4
−2
0
2
4x 10−3
Vyi,j [km/sec]
Time [orbits]
Vy2,1
Vy3,2
Vy4,3
0123456
−2
−1
0
1
2x 10−3
Vzi,j [km/sec]
Time [orbits]
Vz2,1
Vz3,2
Vz4,3
Fig. 3 Relative velocity components in the inertial frame, with an impulsive correction, become bounded.
−3 −2 −1 0 1 2 3 4
−3
−2
−1
0
1
2
3
Yi,j [km]
Xi,j [km]
−3 −2 −1 0 1 2 3 4
−0.03
−0.02
−0.01
0
0.01
0.02
0.03
Zi,j [km]
Xi,j [km]
−3 −2 −1 0 1 2 3
−0.03
−0.02
−0.01
0
0.01
0.02
0.03
Zi,j [km]
Yi,j [km]
−5
0
5
−5
0
5
−0.04
−0.02
0
0.02
0.04
Xi,j [km]
Yi,j [km]
Zi,j [km]
S2,1
S3,2
S4,3
Fig. 4 Relative motion in inertial coordinates exhibits bounded orbits following the formationkeeping maneuver, performed after 1.25 orbits.
1274 BEIGELMAN AND GURFIL
Downloaded by TECHNION - ISRAEL INST OF TECH on December 19, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.34266
A. Calculation of the Required Correction of Semimajor Axis and
Eccentricity
In the presence of J2and drag, the relative orbital elements of any
Siare determined by the GVEs given in Sec. II. The perturbed
dynamics of these elements can be written in the general form:
_
F(68)
where FF1;...;F
6Tis a vector-valued function F:R6!R6.If
the initial orbital elements of Siare different from those of Sj, and in
addition there exists a difference KDin the ballistic coefcients,
then linearizing Eq. (68) about ientails [8]
_
i;j Qi;j qKD(69)
where
QfQlmg@F=@i2R66
and
qq1;...;q
6T@F=@KDi2R6
The entries of Qand qpertinent for the subsequent calculations are
given by [8]
Q41 @F4
@a 21
4J2Re
p2n
acos I(70)
Q42 @F4
@e 6J2R2
e
nae
p3cos I(71)
Q51 @F5
@a Q41
5cos2I1
2 cos I(72)
Q52 @F5
@e Q42
5cos2I1
2 cos I(73)
Q61 @F6
@a 3n
2a17
4J2
1e2
pRe
p2
3cos2I1(74)
Q62 @F6
@e 9
4J2nRe
p2
3cos2I1e
1e25
2
(75)
and
q4@F4
@KD0;q
5@F5
@KD0;q
6@F6
@KD0(76)
To control the orbital-element deviations, impulsive velocity
corrections must be applied. An impulsive velocity correction
causes an instantaneous change of the orbital elements d. The
relationship between the velocity change and the element change
is given by GVEs (3). The velocity correction is chosen to satisfy
[8]:
_
i;j _
i;j;_
M_
!i;j _
M_
!
i;j (77)
When a drag difference exists,
i;j constitutes an initial condition,
and the deviation changes according to Eq. (69). Thus, the desired
drift rates are set to values that will compensate for the effect of
drag differences until next correction.
By combining Eqs. (3) and (69) the following two conditions for
the desired correction values of semimajor axis and eccentricity are
obtained [8]:
L1
i;j q4KD_
i;j (78)
L2
i;j q5q6KD_
M_
!
i;j (79)
where the vector L1is the fourth row of Q,L2is the sum of the fth
and sixth rows of Q, and qiare the elements of q. This procedure
results in
Q41a
i;j Q42e
i;j Q43I
i;j _
i;j (80)
Q51 S61a
i;j Q52 Q62 e
i;j Q53 Q63 I
i;j _
M_
!
i;j
(81)
Equations (81) are two linear equations for the three desired relative
orbital elements a
i;j;e
i;j;I
i;j. This means that there is a single
degree-of-freedom. Because a
i;j and e
i;j depend upon the ballistic
coefcient and I
i;j is unaffected by either J2or by KD, we will
assume that I
i;j 0. This yields two linear equations with two
unknowns, a
i;j and e
i;j, for which the solution is
e
i;j _
M_
!
i;j Q51 Q61_
i;j=Q41
Q52 Q62Q42 =Q41Q51 Q61(82a)
a
i;j
i;j Q42e
i;j
Q41
(82b)
Finally, the optimal velocity correction vis calculated using
Eq. (47) and is given by [compare with Eq. (54)]:
vAa
Ae
a
1;2a
1;2
a
2;3a
2;3
.
.
.
a
N;N1a
2;3
e
1;2e
1;2
e
2;3e
1;2
.
.
.
e
N;N1e
N;N1
2
6
6
6
6
6
6
6
6
6
6
6
6
4
3
7
7
7
7
7
7
7
7
7
7
7
7
5
Aa
Ae
a1;2
a2;3
.
.
.
aN;N1
e1;2
e2;3
.
.
.
eN;N1
2
6
6
6
6
6
6
6
6
6
6
6
6
4
3
7
7
7
7
7
7
7
7
7
7
7
7
5
Aa
Aeb(83)
where Aaand Aeare matrices dened in Eqs. (4851), and the vector
bis a superposition of the precorrected relative orbital elements and
the desired corrections in semimajor axis and eccentricity given by
Eq. (82).
B. Calculation of the Desired Drifts
The desired drifts immediately following the velocity correction,
_
i;j and _
M_
!
i;j, are chosen to be in the opposite direction of the
expected drift due to drag that will be accumulated until the next
correction. These values will be used in Eq. (82). The velocity
correction will yield initial drift rates that are opposite to the drift
due to drag. These drifts can be found using a linear approximation,
valid for a short time interval Tbetween two subsequent corrections
[8]:
_
i;j 1
2Q41F1Q42 F2KD
KD
T(84)
_
M_
!
i;j 1
2Q51 Q61F1Q52 Q62 F2KD
KD
T
(85)
BEIGELMAN AND GURFIL 1275
Downloaded by TECHNION - ISRAEL INST OF TECH on December 19, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.34266
C. Minimization of the Cross-Coupling Effect
A well-known difculty in controlling spacecraft orbits using
impulsive velocity corrections is the cross-coupling between orbital
elements. In formation-ying problems, this effect renders nding
conditions for bounded relative motion quite difcult.
Semimajor axis and eccentricity are corrected by applying in-
plane velocity corrections Vtand Vn. These velocity corrections
in turn cause undesirable changes in Mand !, creating an in-plane
separation M!i;j each time the velocity correction is performed.
However, the orbital elements depend upon true anomaly. This fact
can be used to mitigate the coupling of orbital-element corrections.
For example, a tangential velocity correction applied at periapsis or
apoapsis will not cause any change in !, and a normal velocity
correction applied at f90 or 270 deg will not inuence M.
We will now derive a method for mitigating the in-plane cross-
coupling. This can be done if the semimajor axis and eccentricity
corrections are performed twice per orbit. The rst correction is
performed at f90 deg and the second is performed at
f270 deg. Let us denote the correction of an orbital element
at these points by Iand II, respectively.
Because the change in semimajor axis and eccentricity due to drag
is slow, we can adopt the following approximation:
aIaII (86)
eIeII (87)
This gives
aI2aI2vI
VI
t2aII2vII
VII
taII (88)
From which it immediately follows that VI
tVII
t(under the
assumption that vIvII vand aIaII a). Similarly,
eI2eI
vIVI
trI
aIVI
n2eII
vII VII
trII
aII VII
neII (89)
yielding VI
nVII
n(assuming that eIeII eand
rIrII r).
The parasitic change in !and in Mdue to the rst velocity
correction is given by
!I2
ev VtI2eVI
n(90a)
MI b
eav 21e2r
pVI
t(90b)
and the unwanted change in !and in Mdue to second velocity
correction is given by
!II 2
ev VII
t2eVII
n(91a)
MII b
eav 21e2r
pVII
t(91b)
Table 2 Initial orbital elements of four spacecraft with different
ballistic coefcients
Sia,km eI, rad , rad !, rad f0, rad KD
S16803 0.0012 0.01 0 0 0 0.0355
S26803 0.0012 0.01 0 0 0 0.03905
S36803 0.0012 0.01 0 0 0 0.036
S46803 0.0012 0.01 0 0 0 0.0396
0246810121416
−0.04
−0.02
0
0.02
0.04
ai,j [km]
Time [orbits]
0246810121416
−4
−2
0
2
4x 10−7
ei,j
Time [orbits]
0246810121416
−4
−2
0
2
4x 10−4
fi,j[rad]
Time [orbits]
f2,1
f3,2
f4,3
e2,1
e3,2
e4,3
a2,1
a3,2
a4,3
Fig. 5 Growth of the mean relative semimajor axis, eccentricity and true anomaly due to different ballistic coefcients, without applying
formationkeeping maneuvers.
1276 BEIGELMAN AND GURFIL
Downloaded by TECHNION - ISRAEL INST OF TECH on December 19, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.34266
0246810121416
−4
−2
0
2
4x 10−4
(ω+M)i,j [rad]
Time [orbits]
(ω+M)2,1
(ω+M)3,2
(ω+M)4,3
0246810121416
−1.5
−1
−0.5
0
0.5
1
1.5 x 10−6
i,j [rad]
Time [orbits]
2,1
3,2
4,3
Fig. 6 In-plane and out-of-plane angular separation due to different ballistic coefcients, without applying a formationkeeping maneuver.
0246810121416
−4
−2
0
2
4
Xi,j [km]
Time [orbits]
X2,1
X3,2
X4,3
0246810121416
−4
−2
0
2
4
Yi,j [km]
Time [orbits]
Y2,1
Y3,2
Y4,3
0246810121416
−0.04
−0.02
0
0.02
0.04
Zi,j [km]
Time [orbits]
Z2,1
Z3,2
Z4,3
Fig. 7 Relative drift of the relative position components due to different ballistic coefcients, without applying a velocity correction.
BEIGELMAN AND GURFIL 1277
Downloaded by TECHNION - ISRAEL INST OF TECH on December 19, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.34266
0246810121416
−1
−0.5
0
0.5
1x 10−3
ai,j [km]
Time [orbits]
a2,1
a3,2
a4,3
0246810121416
−1
−0.5
0
0.5
1x 10−8
ei,j
Time [orbits]
e2,1
e3,2
e4,3
0246810121416
−2
0
2x 10−4
fi,j [rad]
Time [orbits]
f2,1
f3,2
f4,3
Fig. 8 Time history of the mean relative values of the semimajor axis and eccentricity, and the true anomaly with different ballistic coefcients, using
optimal velocity corrections.
0 2 4 6 8 10121416
−1.5
−1
−0.5
0
0.5
1x 10−7
(ω+M)i,j [rad]
Time [orbits]
(ω+M)2,1
(ω+M)3,2
(ω+M)4,3
0 2 4 6 8 10121416
−3
−2
−1
0
1
2
3
4x 10−10
i,j [rad]
Time [orbits]
2,1
3,2
4,3
Fig. 9 Time history of the in-plane and out-of-plane angular separation with different ballistic coefcients using periodic velocity corrections. The
angular drifts are minuscule.
1278 BEIGELMAN AND GURFIL
Downloaded by TECHNION - ISRAEL INST OF TECH on December 19, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.34266
0246810121416
−2
−1
0
1
2x 10−3
Vxi,j [m/sec]
Time [orbits]
Vx2,1
Vx3,2
Vx4,3
0246810121416
−2
−1
0
1
2x 10−3
Vyi,j [m/sec]
Time [orbits]
Vy2,1
Vy3,2
Vy4,3
0246810121416
−1
−0.5
0
0.5
1x 10−4
Vzi,j [m/sec]
Time [orbits]
Vz2,1
Vz3,2
Vz4,3
Fig. 10 Relative velocity components with different ballistic coefcients and periodic velocity corrections.
0246810121416
−2
0
2
4
Xi,j [m]
Time [orbits]
X2,1
X3,2
X4,3
0246810121416
−2
−1
0
1
2
Yi,j [m]
Time [orbits]
Y2,1
Y3,2
Y4,3
0246810121416
−0.02
−0.01
0
0.01
0.02
Zi,j [m]
Time [orbits]
Z2,1
Z3,2
Z4,3
Fig. 11 The position components in inertial coordinates with different ballistic coefcients under periodic velocity corrections, optimally calculated to
mitigate the effect of differential perturbations.
BEIGELMAN AND GURFIL 1279
Downloaded by TECHNION - ISRAEL INST OF TECH on December 19, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.34266
By combining Eqs. (90) with Eqs. (91), we get the total in-plane
deviation during an orbital period:
!I!IIMIMII 0(92)
which means that if we perform in-plane corrections at these points, a
second in-plane change will cancel out the rst parasitic in-plane
change.
D. Simulation
First, we will simulate the uncontrolled motion of the spacecraft in
a 425-km low Earth orbit. The differences in ballistic coefcients
between the spacecraft are about 10%. The initial orbital elements
and ballistic coefcients are summarized in Table 2. As can be seen in
Figs. 5 and 6, different ballistic coefcients cause different drifts in
semimajor axis and eccentricity, which in turn induces secular differ-
ences in the out-of-plane and in-plane angular motions due to the J2
effect. Figure 7 depicts the relative position components in inertial
coordinates. The difference in ballistic coefcients increases the rela-
tive position components so that the average distance between the
formation-ying spacecraft amount to about 2.5 km after 16 orbits.
Next, we demonstrate that the proposed formationkeeping method
[Eq. (83)] considerably reduces the relative drift. The orbital elements
are corrected twice per orbit, at f90 and 270 deg, thus minimizing
the effect of cross-coupling between the orbital elements. As shown in
Figs. 8 and 9, the periodic impulsive corrections almost nullify the
relative drift in the relative orbital elements. Figures 10 and 11 show
the time history of the relative velocity and position components,
respectively, in inertial coordinates. The discontinuities in the relative
velocity components are a result of the impulsive velocity corrections
that are performed twice per orbit. Figure 12 shows the three-
dimensional relative orbits and the projections of the relative orbits
onto the XY,XZ, and YZplanes in inertial coordinates. Figure 13
depicts the distance between spacecraft. The drift is not zero, but it is
much smaller than the uncorrected case (about three orders of
magnitude, relative to the uncorrected case). A small in-plane drift
remains because the cross-coupling effect causes a small residual drift
in relative position. The actual in-plane drift is not identically zero
because of the assumption we made that the change in semimajor axis
and eccentricity due to drag is slow. The maximum separation is
between S4and S3and is about 2.5 m after 16 orbits (one day).
E. Fuel Consumption
The fuel required for formationkeeping is related to the total veloc-
itychangeby the rocket equation(35).The total velocity changecanbe
calculated by either Eq. (33) or Eq. (34), depending on the type of
propulsion system. For a mission lifetime of ve years, the total veloc-
ity corrections of each spacecraft, calculated according to Eq. (34), are
vtot119:89 m=s;vtot214:78 m=s
vtot314:97 m=s;vtot420:26 m=s(93)
−2 −1 0 1 2 3
−2
−1
0
1
2
Yi,j [m]
X i,j [m]
−2 −1 0 1 2 3
− 0.02
− 0.01
0
0.01
0.02
Zi,j [m]
Xi,j [m]
−2 −1 0 1 2
− 0.02
− 0.01
0
0.01
0.02
Zi,j [m]
Yi,j [m]
−5
0
5
−2
0
2
− 0.02
0
0.02
Xi,j [m]
Yi,j [m]
Zi,j [m]
S2,1
S3,2
S4,3
Fig. 12 Relative motion between spacecraft with different ballistic coefcients in inertial coordinates, using periodic velocity corrections. The drift is
reduced by three orders of magnitude compared with the uncontrolled case.
1280 BEIGELMAN AND GURFIL
Downloaded by TECHNION - ISRAEL INST OF TECH on December 19, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.34266
and the total vis about 70 m=s. The actual propellent consumption
can be calculated according to the rocket equation (35). Taking
Isp 220 s and m0500 kg, the required fuel mass for each
spacecraft is as follows:
m14:59 kg;m23:41 kg
m33:46 kg;m44:67 kg (94)
We thus see that the propellent mass is properly balanced among the
formation spacecraft. For comparison, the total ve-year velocity
corrections calculated according to Eq. (33) are
vtot123:12 m=s;vtot217:16 m=s
vtot317:45 m=s;vtot423:41 m=s(95)
and the total velocity correction is about 81 m=s.
VII. Conclusions
We developed a generic fuel-balanced impulsive maneuver
scheme for multiple spacecraft formationkeeping using relative-
orbital-element corrections. The spacecraft formation was modeled
as a directed graph, an approach that accommodates any number of
spacecraft. More important, formulating the problem of
formationkeeping in terms of relative-orbital-element corrections
left the nal values of the controlled elements unconstrained. This
freedom was used to nd optimal impulsive maneuvers, in the sense
of minimizing the l2-norm, for formation initialization and control in
the presence of perturbations.
The main conclusions are as follows:
1) A controlled group of spacecraft will form a pattern in the
position space that corresponds to the optimal velocity correction.
2) Leaving the reference orbit free results in a globally optimal
velocity correction.
3) This velocity correction can be analytically computed using a
simple least-squares procedure.
Because the formationkeeping formalism is formulated in terms of
relative-orbital-element corrections, one can choose which orbital
elements to control. The nal differences in the controlled elements
can be constrained to minimize the effect of orbital perturbations and
cross-coupling.
The newly developed algorithm was illustrated by a simulation of
a formation of four spacecraft ying in low Earth orbits subject to J2
and drag perturbations. Two different scenarios were examined.
The rst is when all spacecraft possess identical ballistic
coefcients, and the second is when the spacecraft have different
ballistic coefcients.
In the rst scenario, the most dominant perturbation is J2. Because
of initialization errors, small initial differences of the semimajor axis,
eccentricity, and inclination lead to secular changes in both the nodal
rate and the mean latitude rate. The formationkeeping algorithm is
then required to cancel theses relative mean drifts. This was achieved
by matching semimajor axis, eccentricity, and inclination.
In the second scenario, a periodic velocity correction should be
applied to preserve the formation, because the corrected orbital
elements deviate from the reference values due to different ballistic
coefcients. We proposed controlling the nodal rate and the mean
latitude rate indirectly, by selecting suitable values of semimajor axis
and eccentricity. These values were selected to cancel future drifts
until the next velocity correction. The simulation results showed that
applying the newly developed formationkeeping reduced the relative
drift by three orders of magnitude.
Appendix A: Effect of Different Spanning Trees on the
Velocity-Correction Vector
According to Cayleys formula, the number of different spanning
trees of the connected graph GNis NN2, where each spanning tree
forms a different labeled graph [22]. In the previous sections, we
treated the spacecraft formation as a leader formation LN, which is
also a labeled graph. The leader of this formation was labeled S1.
This means that for a formation of Nspacecraft, we can form NN2
different Amatrices and bvectors. The analysis in this paper was
performed for a formation represented by the particular spanning
tree
SN!SN1!!S1
We will now examine how different spanning trees affect the
velocity correction required from each spacecraft and the total
velocity correction. For simplicity, we will study only semimajor-
axis corrections. To that end, consider all 16 possible spanning trees
for a formation of four spacecraft, as shown in Fig. A1. We
arbitrarily choose four different spanning trees: L4
1,L4
7,L4
13, and
L4
16. These trees are shown in Fig. A1. The initial values of
semimajor axes for this formation are as follows:
a142;160 km;a
242;167 km;
a342;170 km;a
442;185 km (A1)
The matrices Aa1,Aa7,Aa13, and Aa16 and the vectors b1,b7,b13 ,
and b16, corresponding to the trees L4
1,L4
7,L4
13, and L4
16 are as
follows:
Aa1
2v1a2
1
2v2a2
2
00
02v2a2
2
2v3a2
3
0
00
2v3a2
3
2v4a2
4
2
6
6
43
7
7
5
;b1
a2;1
a3;2
a4;3
2
43
5
(A2)
Aa7
2v1a2
1
2v2a2
2
00
2v1a2
1
02v3a2
3
0
00
2v3a2
3
2v4a2
4
2
6
6
43
7
7
5
;b7
a2;1
a3;1
a4;3
2
43
5
(A3)
Aa13
2v1a2
1
2v2a2
2
00
02v2a2
2
2v3a2
3
0
02v2a2
2
02v4a2
4
2
6
6
43
7
7
5
;b13
a2;1
a3;2
a4;2
2
43
5
(A4)
0246810121416
0
0.5
1
1.5
2
2.5
Relative distance [m]
Time [orbits]
d2,1
d3,2
d4,3
Fig. 13 Relative distance between spacecraft with different ballistic
coefcients shows that a maximum separation of only 2.5 m after 16
orbits.
BEIGELMAN AND GURFIL 1281
Downloaded by TECHNION - ISRAEL INST OF TECH on December 19, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.34266
Aa16
2v1a2
1
2v2a2
2
00
2v1a2
1
02v3a2
3
0
2v1a2
1
00
2v4a2
4
2
6
6
43
7
7
5
;b16
a2;1
a3;1
a4;1
2
43
5
(A5)
Substitution into Eq. (47) yields the optimal velocity-correction
vector, which is equal for the four different trees already described, as
is the nal value of the semimajor axis of this formation (without
perturbations):
v?
0:3826
0:1273
0:0180
0:5285
2
6
6
43
7
7
5
m=s;a
a?42;170:49 km (A6)
This example shows that different spanning trees do not affect the
control vector and the required velocity corrections of each
spacecraft. This is because different spanning trees are merely a
different parametrization of the same physical values of orbital
elements. The nal value of the orbital element is an optimal value
for given initial conditions. Choosing different spanning trees does
not change these initial values and the concomitant required
corrections .
Acknowledgments
This research was partially supported by the Israel Ministry of
Science and Technology Infrastructure Program. The authors owe a
debt of gratitude to Moshe Guelman and David Mishne for providing
important insights.
References
[1] Clohessy, W., and Wiltshire, R., Terminal Guidance System for
Satellite Rendezvous,Journal of the Astronautical Sciences, Vol. 27,
No. 9, 1960, pp. 653678.
[2] Karlgaard, C. D., and Lutze, F. H., Second-Order Relative Motion
Equations,Journal of Guidance, Control, and Dynamics, Vol. 26,
No. 1, Jan. 2003, pp. 4149.
[3] Gurl, P., and Kasdin, N. J., Nonlinear Modeling of Spacecraft
Relative Motion in the Conguration Space,Journal of Guidance,
Control, and Dynamics, Vol. 27, No. 1, 2004, pp. 154157.
doi:10.2514/1.9343
[4] Chichka, D., Satellite Clusters with Constant Apparent Distribution,
Journal of Guidance, Control, and Dynamics, Vol. 24, No. 1, 2001,
pp. 117122.
[5] Alfriend, K., Schaub, H., and Gim, D., Gravitational Pertubations,
Nonlinearity and Circular Orbit Assumption Effects on Formation
Flying Control Strategies,23rd AAS Guidance and Control Confer-
ence, American Astronautical Society Paper AAS-00-012, 2000.
[6] Schaub, H., and Alfriend, T., J2Invariant Relative Orbits for
Spacecraft Formations,Celestial Mechanics and Dynamical
Astronomy, Vol. 79, No. 2, 2001, pp. 7795.
doi:10.1023/A:1011161811472
[7] Schaub, H., Vadali, S. R., and Alfriend, K. T., Spacecraft Formation
Flying Control Using Mean Orbital Elements,Journal of the
Astronautical Sciences, Vol. 48, No. 1, 2000, pp. 6987.
[8] Mishne, D., Formation Control of Satellites Subject to Drag Variations
and J2Perturbations,Journal of Guidance, Control, and Dynamics,
Vol. 27, No. 4, 2004, pp. 685692.
doi:10.2514/1.11156
4
1
2
3 3
1
2
4 4
1
3
2 2
1
3
42
1
4
3 3
1
4
2
1
3
4
2 1
2
4
31
3
2
41
4
2
31
2
3
41
4
3
2
1
2
3
4
1
3
2
4
1
4
3
2
4
1
3
2
4
1
L4
2
L
4
3
L
4
4
L
4
5
L
4
6
L
4
7
L
4
8
L
4
9
L
4
10
L
4
11
L
4
12
L
4
13
L
4
14
L
4
15
L
4
16
L
Fig. A1 Sixteen different spanning trees for a formation of four spacecraft.
1282 BEIGELMAN AND GURFIL
Downloaded by TECHNION - ISRAEL INST OF TECH on December 19, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.34266
[9] Schaub, H., and Alfriend, K., Hybrid Cartesian and Orbit Elements
Feedback Law for Formation Flying Spacecraft,Journal of
Guidance, Control, and Dynamics, Vol. 25, No. 2, Mar.Apr. 2002,
pp. 387393.
[10] Gurl, P., Relative Motion Between Elliptic Orbits: Generalized
Boundedness Conditions and Optimal Formationkeeping,Journal of
Guidance, Control, and Dynamics, Vol. 28, No. 4, July 2005, pp. 761
767.
doi:10.2514/1.9439
[11] Hughes, S. P., Cooley, D. S., and Guzman, J. J., A Direct Approach for
Minimum Fuel Maneuvers of Distributed Spacecraft in Multiple Flight
Regimes,Advances in the Astronautical Sciences, Vol. 120, No. 1,
2005, pp. 869898.
[12] Schaub, H., Incorporating Secular Drifts into the Orbit Element
Difference Description of Relative Orbits,Advances in the
Astronautical Sciences, Vol. 114, 2003, pp. 239257.
[13] Battin, R. H., An Introduction to the Mathematics and Methods of
Astrodynamics, AIAA, Reston, VA, 1999, pp. 488489, Chap. 10.
[14] Tillerson, M., Breger, L., and How, J. P., Distributed Coordination and
Control of Formation Flying Spacecraft,Proceedings of the American
Control Conference, Vol. 2, American Automatic Control Council,
Evanston, IL, 2003, pp. 17401745.
[15] Ross, I. M., King, J., and Fahroo, F., Designing Optimal Spacecraft
Formations,AIAA/AAS Astrodynamics Conference, AIAA
Paper 2002-4635, Aug. 2002.
[16] Yang, G., Yang, Q., Kapila, V., Palmer, D., and Vaidyanathan, R.,
Fuel Optimal Manoeuvres for Multiple Spacecraft Formation
Reconguration Using Multi-Agent Optimization,International
Journal of Robust and Nonlinear Control, Vol. 12, No. 3, 2002,
pp. 243283.
doi:10.1002/rnc.684
[17] Vadali, S. R., Vaddi, S. S., and Alfriend, K. T., An Intelligent
Control Concept for Formation Flying Satellites,International
Journal of Robust and Nonlinear Control, Vol. 12, No. 3, Mar. 2002,
pp. 97115.
doi:10.1002/rnc.678
[18] Diestel, R., Graph Theory, SpringerVerlag, New York, 2005, pp. 1
30.
[19] Mesbahi, M., and Hadaegh, F., Formation Flying Control of Multiple
Spacecraft via Graphs, Matrix Inequalities, and Switching,Journal
of Guidance, Control, and Dynamics, Vol. 24, No. 2, 2001, pp. 369
377.
[20] Vallado, D. A., Fundamentals of Astrodynamics and Applications,
2nd ed., 2001, pp. 135235.
[21] Ross, I. M., Space Trajectory Optimization and L1-Optimal Control
Problems,Modern Astrodynamics, edited by P. Gurl, Elsevier,
Amsterdam, 2006, pp. 155188.
[22] Even, S., Graph Algorithms, Computer Science Press, New York, 1979,
pp. 150154.
BEIGELMAN AND GURFIL 1283
Downloaded by TECHNION - ISRAEL INST OF TECH on December 19, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.34266
... It is worth mentioning that, for those cluster establishment missions allowing some flexibility to determine the target (final) states, the existing few works are devoted to either a decoupled control scheme [22] or a static l 2 -norm optimization [23], which, however, actually corresponds to the energy-optimal case. The underlying methodology of this Note, on the contrary, relies on an indirect approach, capable of identifying the target states as a natural output of the optimization process. ...
... In cluster flight missions, an efficient means to reduce the amount of propellant required to counter the undesirable perturbation effects is the proper initialization of the spacecraft relative orbits [23]. As shown in previous works [17,18], the necessary condition eliminating the mutual separation among the cluster is to match the secular drift rate of f a; e; i; Ω R g resulting from an SDSOGF. ...
... In fact, Eqs. (21)(22)(23) can be collected into one equality by eliminating ξ Δ p , viz., ...
... A basic idea of the filtering approach is to estimate the angular rate dependent on angular dynamics of a rigid-body spacecraft and correct prediction errors by measurements from a star tracker [11][12][13]. However, the angular momentum and the inertia matrix in the dynamics model may contain parameter uncertainties due to difficulties of accurate modeling, fuel consumptions, and external disturbances, eventually leading to the performance deterioration or instability of angular rate estimation [14,15]. As for the derivative approach, the angular rate can be derived by differentiating the attitude given by a star tracker [16]. ...
... For the variable angular rate, it is reasonable to assume that the angular rate is constant in the short time interval between two adjacent star spots extracted in the RS mode. Based on Equation (8) and Equation (14) the angular rate is determined by the least square approach. ...
Article
Full-text available
Angular rate is a piece of useful information for the attitude control of a spacecraft. The star tracker as a space optical sensor can be used to measure the angular rate of a spacecraft. In this paper, a novel approach is proposed to improve the measurement accuracy of the angular rate during spacecraft rotation. The electronic rolling shutter (RS) imaging mode of the complementary metal-oxide semiconductor (CMOS) image sensor in a star tracker is applied to obtain much higher sampling frequency for reducing the change of the angular rate between the sampling interval. The optic flow vector on the imaging plane is approximated within the second order using three successive star images to reflect the nonlinear effect from the variable angular rate. The experiment is performed to demonstrate the advantage of the new approach for variable angular rate measurement.
... Specifically, Hamel et al. 10 formulated an optimal feedback control law for J2-perturbed earth-orbiting formation flying spacecraft which uses relative orbital elements as the states. Furthermore, Beigelman and Gurfil 11 developed an optimal fuel-balanced impulsive formation keeping control under J2 and atmospheric drag perturbations. However, in all these strategies, the dynamical models were used to formulate control laws and not directly used in guidance formulations. ...
Conference Paper
A novel guidance law that utilizes optimal control theory in combination with linear time-varying state-transition matrices to calculate the optimal path for a chaser spacecraft from its initial point to a desired final point and a desired amount of time is proposed in this paper. In addition, this paper presents a new method that allows for the closed-loop testing of guidance, navigation and control systems with realistic software models of GPS receivers and RADAR systems for on-board relative navigation. Specifically, utilizing the signal-to-noise ratio of various sensors, the errors that arise due hardware noise and/or environmental factors can be modeled accurately based on the communication link. The guidance law was tested using the newly developed software-in-the-loop test-bed for a far-range formation flight scenario around the non-cooperative Alouette-2 spacecraft as well as on a highly elliptical orbit.
... Meanwhile, et NW and et SE fire twice at ascending nodes (l NW1 , l NW2 ) and descending nodes (l SE1 , l SE2 ) respectively in each control period to maximize the stationkeeping efficiency. Thus, the firing right ascensions of the electric thrusters are depicted in Eq. (16) (Beigelman and Gurfil, 2008) ...
Article
Full-text available
This paper proposes a novel electric thruster configuration design optimization scheme to implement efficient station-keeping operations for geostationary satellites. In this work, the electric thrusters are equipped with 4-DOF robotic manipulators to provide flexible control thrust in space, which consumes much less propellant than the conventional station-keeping process using fixed-position electric thrusters. To effectively determine the firing right ascensions and firing durations of each electric thruster during the station-keeping process, a novel station-keeping control strategy is developed in view of the low-thrust feature of electric thrusters and the flexible pointing capacity of robotic manipulators. A low-thrust station-keeping period includes one two-day orbit determination period and six two-day control periods, where the initial orbital elements are determined in the orbit determination period, and the electric thrusters fire to compensate the perturbations in the control periods. Based on the control strategy, the electric thruster configuration design optimization problem is formulated to minimize the total propellant consumption via optimizing the robotic manipulator joint swivel angles and firing durations simultaneously, subject to several practical engineering constraints including the station-keeping accuracy, angular momentum dumping, collision between electric thruster and solar array, and electric thruster plume. Finally, the proposed electric thruster configuration design optimization scheme is applied to real-world station-keeping problems. The simulation results show that the geostationary satellites can be successfully maintained in an assigned station-keeping window using the robotic manipulators based low-thrust station-keeping control strategy. Moreover, the optimized electric thruster configuration with robotic manipulators can save 31.58% propellant consumption compared with that of the conventional configuration owing to the favorable flexibility of control thrust, which demonstrates the effectiveness and practicality of this work.
... Schaub et al. [2] proposed using differential mean orbital elements as feedback states for determining impulsive orbit corrections. Beigelman and Gurfil [3] developed an optimal fuel-balanced impulsive control algorithm. Other impulsive control approaches can be found in Refs. ...
Article
Full-text available
Using differential solar radiation pressure (DSRP) for satellite formation keeping is not a new idea. However, there is no exiting design methodology of DSRP-based controllers for lunar satellite formation keeping. Lunar orbits are different from low Earth orbits, because the third-body effects are dominant. This paper develops a DSRP-based formation keeping controller, utilizing mean orbital element feedback, for regulating relative distances between satellites in lunar orbits. It is shown that the most effective way to mitigate the inter-satellite drift, is to adjust the semimajor axes. This is achieved by modifying the cross-sectional areas of the satellites. The stability of the closed-loop system is proven based on finite-time stability theory. Numerical simulation results illustrate that the new DSRP-based controller is able to arrest the relative distance drift in lunar orbits for several years.
Article
A novel guidance law that utilizes optimal control theory in combination with linear time-varying state-transition matrices to calculate the optimal path for a chaser spacecraft from its initial point to a desired final point and a desired amount of time is proposed in this paper. The optimal guidance law is formulated using more accurate second-order integral form of Gauss Variational Equations to map control accelerations to the dynamics in terms of relative orbital elements. When comparing the second-order form against the traditional Gauss Variational Equations, the guidance law results in a 15% reduction in control demand. In addition, this paper presents a new method that allows for the closed-loop testing of guidance, navigation and control systems with realistic software models of GPS receivers, RADAR systems for on-board relative navigation, and ground-station up-links of the target states. Specifically, utilizing the signal-to-noise ratio of various sensors, the errors that arise due to hardware noise and/or environmental factors can be modeled accurately based on the communication link. The guidance law is tested using the newly developed software-in-the-loop test-bed for a far-range formation flight scenario around the non-cooperative Alouette-2 spacecraft as well as on an arbitrary defined highly elliptical orbit.
Article
Full-text available
Based on Gauss’s variational equations (GVEs), the impulsive orbital-element corrections are investigated in orbit transfer problems. Both single impulse and multiple impulses are considered for the first-order and second-order GVEs. For the single impulse, a nonlinear least-squares iteration method for the minimum orbit error is provided to simultaneously solve for the impulse vector and the impulse position. For multiple impulses, a minimum-norm method for the energy cost is proposed to solve the three-impulse and two-impulse corrections for the in-plane and out-of-plane orbital elements, respectively. The impulse positions are analytically derived, and the impulse vectors are obtained by the minimum-norm method. Numerical examples are provided to verify the proposed single-impulse least-squares and multiple-impulse minimum-norm methods. The results show that the nonlinear solutions using the second-order GVEs are more accurate than those using the first-order GVEs.
Article
A robust trajectory optimization approach for rendezvous and proximity operations in perturbed elliptical orbits with uncertainties is proposed. For robust trajectory design, a new discrete-time linear dynamics model describing the relative equations of motion for powered flight in perturbed elliptical orbits is developed. The linear dynamics model is used to formulate a stochastic trajectory optimization problem that takes into account navigation sensor errors, maneuver execution errors, and trajectory and control dispersion. The objective of the stochastic optimization problem is to minimize the sum of the expected maneuver and the maneuver dispersion subject to a 3-σ constraint on the final state dispersion. A genetic algorithm is used to solve the stochastic trajectory optimization problem, and a series of examples demonstrate that the stochastic maneuver solution can be significantly better than the solution corresponding deterministic trajectory optimization problem. Nonlinear Monte Carlo simulations are used to validate the results.
Article
Full-text available
Based on the concept of orbital commensurability, necessary and sufficient conditions are presented for bounded relative motion between any two spacecraft flying on elliptic Keplerian orbits. The proposed approach does not involve any simplifying assumptions regarding the relative dynamics but rather treats the general, nonlinear, eccentric relative motion problem. The methodology presented alleviates the difficulty in computing corrections to the linear equations of motion to account for nonlinearities and eccentricities. Instead of dealing with the local relative motion problem, the global relative motion problem is addressed by transforming the orbital resonance requirement into an energy-matching condition. The newly developed setup is then utilized to derive an optimal single-impulse formationkeeping maneuver based on relative state variables. The orbital elements interpretation of the optimal formationkeeping maneuver is also discussed.
Article
Hill's equations have been used in most formation flying studies to determine relative motion orbits and control strategies. Hill's equations assume the Chief satellite orbit is circular, the Earth is spherically symmetric and the nonlinear terms in the relative motion variables can be neglected. This paper presents an approach for determining the effect of these assumptions on the fuel consumption for establishing and maintaining a relative motion orbit. Initial results on the errors in predicting the relative motion using Hill's equations are presented.
Article
An approximate solution of second-order relative motion equations is presented. The equations of motion for a Keplerian orbit in spherical coordinates are expanded in Taylor series form using reference conditions consistent with that of a circular orbit. Only terms that are linear or quadratic in state variables are kept in the expansion. The method of multiple scales is employed to obtain an approximate solution of the resulting nonlinear differential equations, which are free of false secular terms. This new solution is compared with the previously known solution of the linear case to show improvement and with numerical integration of the quadratic differential equation to understand the error incurred by the approximation. In all cases, the comparison is made by computing the difference of the approximate state (analytical or numerical) from numerical integration of the full nonlinear Keplerian equations of motion. The results of two test cases show two orders of magnitude improvement in the second-order analytical solution compared with the previous linear solution over one period of the reference orbit.
Article
Abstract Two nonlinear feedback control laws are presented to reestablish a desired J2 invariant relative orbit. Since it is convenient to describe the relative orbit of a deputy satellite with respect to a chief satellite in terms of mean orbit element dierences, and because the conditions for a relative orbit being J2 invariant are expressed in terms of mean orbit elements, the first control law feeds back errors in terms of mean orbit elements. Dealing with mean orbit elements has the advantage that short period oscillations are not perceived as tracking errors; rather, only the long term tracking errors are compensated,for. The second control law feeds back traditional cartesian position and velocity tracking errors. For both of the control laws, the desired orbit is computed using mean orbit elements. A numerical study compares and contrasts the two feedback laws.
Article
The fundamental mathematical principles of celestial mechanics and space navigation are presented in an introduction for students of aerospace engineering and science. Chapters are devoted to hypergeometric functions and elliptic integrals, basic topics in analytical dynamics, the two-body problem, two-body orbits and the initial-value problem, solving the Kepler equation, the two-body orbital boundary-value problem, and solving Lambert's problem. Also examined are non-Keplerian motion, patched-conic orbits and perturbation methods, variation of parameters, two-body orbital transfer, numerical integration of differential equations, the celestial position fix, and space navigation. Diagrams, graphs, and sample problems are provided.
Article
The issue of controlling the relative position of satellites in formation has acquired increasing attention in recent years. Special challenge is the control of a low-Earth-orbit formation, where differences in Earth oblateness effects and different atmospheric drag cause a drift of the relative position. A method is developed to compensate for the secular combined effects of first-order gravitational perturbations (J(2)) and atmospheric drag perturbations. The compensation is performed by impulsive velocity corrections. The velocity correction vector and the timing of the correction are determined such that the secular drifts of the relative nodal rate and of the relative mean latitude rate are set to values that cancel out future drifts due to drag for the next time interval. In addition, an optimality condition is developed such that this correction velocity is minimized. After the correction, the orbital element deviations are not set to zero, but rather to small acceptable values for which the in-plane and the out-of-plane relative drifts are small and bounded. The algorithms for the determination of the velocity correction are developed, and a numerical example is presented. The effects of measurement noise and drag uncertainty are discussed as well.
Article
The problem of creating a cluster of satellites such that the arrangement maintains its apparent relative formation as viewed from the planetary surface, without active station keeping, is considered. It is shown that it is possible to create a cluster with nearly constant circular form, but that rotates about its center. The amount of variation from the perfect circle is characterized in terms of the eccentricity of the satellite orbits. Because of the nature of the cluster, it is possible to put a large number of satellites into a cluster, with no danger of satellite collisions. An illustrative example is included.