Survey of Maneuvering Target Tracking. Part II Motion Models of Ballistic and Space Targets-KvL.pdf

(2404 KB) Pobierz
641081518 UNPDF
I. INTRODUCTION
Survey of Maneuvering
Target Tracking.
Part II: Motion Models of
Ballistic and Space Targets
A survey of dynamic models used in maneuvering
target tracking has been reported in [64]. It, however,
does not cover motion models used for tracking
ballistic and space targets (BT), namely, ballistic
missiles, decoys, debris, satellites, projectiles, etc.
The primary reason for this omission is that these
models possess many distinctive features that differ
vastly from those covered in [64]. To supplement
[64], a survey of these models is presented here. To
our knowledge, such a survey is not available in the
literature.
Overall, a BT has a less uncertain motion than
many other types of powered vehicles, such as
maneuvering aircraft or agile missiles: Most BTs
follow a flight path that is largely predetermined
by the performance characteristics specific for a
target type, hence the name ballistic targets, although
some more advanced ballistic missiles can undergo
small maneuvers usually for retargeting. However,
this does not mean that the motion of a foreign BT
can be determined accurately. In fact, tracking a
foreignBTislikelytobeplaguedwithavarietyof
uncertainties, including those concerning trajectory
loft or depression, thrust profile management, target
weight, propellant specific impulse, sensor bias, and
atmospheric parameters. Many of these uncertainties
stem from the uncertainty in the target type, the
principal uncertainty in modeling the motion of a
foreign BT for the tracking purpose.
The entire trajectory of a BT is commonly
divided into three basic phases: boost (and possibly
a post-boost phase), ballistic (also known as coast),
and reentry. The boost phase is the powered,
endo-atmospheric flight, which lasts from launch to
thrust cutoff or burnout. It is followed by the ballistic
phase, which is an exo-atmospheric, free-flight motion
for most BTs, continuing until the atmosphere is
reached again. The atmospheric reentry begins when
the atmospheric drag becomes considerable and
endures until impact.
For the tracking purpose, only the most substantial
forces that may act on a BT are considered: thrust,
aerodynamic forces (most notably, atmospheric
drag and possibly lift), and gravity. Not all of these
forces are present at a level that significantly affects
the motion of a BT in all regimes of the trajectory.
The boost phase is characterized by a large thrust,
which in the case of rocket staging is subject to
abrupt, jump-wise changes. The effects of the drag
and gravity are also essential in this phase. After the
boost, drag is no longer present and thrust vanishes or
drops to a very low level. During the exo-atmospheric
ballistic phase the motion is governed essentially by
the gravity only. Still, small retargeting maneuvers
are possible. The reentry phase features a rapid
drag-induced deceleration with possible lateral
X. RONG LI, Fellow, IEEE
VESSELIN P. JILKOV, Member, IEEE
University of New Orleans
This paper is the second part in a series that provides a
comprehensive survey of maneuvering target tracking without
addressing the so-called measurement-origin uncertainty. It
surveys motion models of ballistic targets used for target tracking.
Models for all three phases (i.e., boost, coast, and reentry) of
motion are covered.
CONTENTS
I. Introduction
II. Ballistic (Coast) Flight
A. Gravity
B. Coordinate Accelerations
C. Coast Motion Models
III. Reentry
A. Aerodynamic Forces
B. Ballistic (Nonmaneuvering) RV Motion
C. Maneuvering RV Motion
D. Motion of Endo-Atmospheric Ballistic Targets
IV. Boost Phase
A. Accelerations
B. Kinematic Models
C. Dynamic Models
D. Example of A Profile-Based Model
V. Integration of BT Motion Models for Entire Trajectory
VI. Concluding Remarks
References
Appendix
Part I: Dynamic Models. IEEE Transactions on Aerospace and Electronic
Systems , 39 , 4 (Oct. 2003), 1333—1364.
Part II: Originally published as Ballistic Target Models. In Proceedings of
the 2001 SPIE Conference on Signal and Data processing of Small Targets ,
vol. 4473, 559—581.
Part III: Measurement Models. In Proceedings of the 2001 SPIE Conference
on Signal and Data Processing of Small Targets , vol. 4473, 423—446.
Part IV: Decision-Based Methods. In Proceedings of the 2002 SPIE
Conference on Signal and Data Processing of Small Targets , vol. 4728,
511—534.
Part V: Multiple-Model Methods. IEEE Transactions on Aerospace and
Electronic Systems , 41 , 4 (Oct. 2005), 1255—1321.
Manuscript received August 4, 2007; revised April 8, 2008; released
for publication August 25, 2008.
IEEE Log No. T-AES/46/1/935930.
Refereeing of this contribution was handled by W. Koch.
This research was supported in part by ARO Grant
W911NF-08-1-0409, ONR-DEPSCoR Grant N00014-09-1-1169,
Project 863 through Grant 2006AA01Z126, and Louisiana BoR
Grant LEQSF (2009-12)-RD-A-25.
Authors’ address: Dept. of Electrical Engineering, University of
New Orleans, New Orleans, LA 70148, E-mail: (xli@uno.edu).
0018-9251/10/$26.00 ° 2010 IEEE
96
IEEE TRANSACTIONS ON AEROSPACE AND ELECTRONIC SYSTEMS VOL. 46, NO. 1 JANUARY 2010
AUTHORIZED LICENSED USE LIMITED TO: IEEE XPLORE. DOWNLOADED ON MAY 13,2010 AT 11:53:33 UTC FROM IEEE XPLORE. RESTRICTIONS APPLY.
641081518.004.png 641081518.005.png
accelerations. So, the significant forces in the different
phases are boost (thrust, gravity, and aerodynamic);
coast (gravity); reentry (aerodynamic and gravity).
In general, the total acceleration vector of a BT, in
an inertial coordinate system (CS) (e.g., the ECI-CS,
see Fig. 2 in the Appendix) can be decomposed as a
vector sum:
a = a T + a A + a G = a T + a D + a L + a G
(1)
same disclaimers as made in the Introduction of
[64] apply to this survey, including those on the
restriction of point targets for temporal behaviors and
the interrelationship between dynamic models and
tracking algorithms. The reader should keep in mind
that results of many extensive studies of BT tracking
have not been published in the open literature. In
other words, much of the BT information, particularly
target-type specific information, is classified and not
open to the general public. As a result, this survey
covers only those dynamic models for BT tracking
in the open literature available to us. The emphasis
of the survey is on missile tracking, although some
models covered are applicable to tracking other
targets.
This paper is a heavily revised and updated version
of [61]. It is organized as follows: Motion models for
the simplest phase, the ballistic flight, are covered in
Section II. This is followed in Section III by a survey
of the models for reentry vehicles. Section IV then
describes models for the most sophisticated, the boost
phase. Integration of motion models for different
phases is the topic of Section V. Concluding remarks
are made in Section VI. The Appendix provides
background information for the most common CSs
used in BT tracking.
where a T , a A ,and a G denote the accelerations
induced by thrust, aerodynamic forces, and gravity,
respectively, and a A consists of accelerations due to
drag, a D ,andlift, a L . In a frame fixed to the Earth, the
relative total acceleration also includes those induced
by the Earth’s rotation.
The entire end-to-end motion of a BT can be
modeled by a “wide-band” dynamic model (e.g.,
nearly constant velocity, acceleration, jerk, or Singer
model) capable of covering all the phases. Most
models developed for the boost phase, the most
sophisticated of all three phases, can serve this
purpose. This is, however, rather crude and not in
common use. Target dynamics during the different
phases are substantially different. A succinct outline
of the main dynamic phases is given in [28]. What is
more natural and rational, as well as common practice,
is to develop different models specific for each phase
that more fully exploit the inherent characteristics of
the portion.
Throughout this paper, let the target position
and velocity vectors be p =[ X , Y , Z ] 0 and v = p =
[ _ X , _ Y , _ Z ] 0 , respectively, in whatever coordinate
system is considered. For example, p =
II. BALLISTIC (COAST) FLIGHT
A. Gravity
¡!
OP in the
(Earth-centered-inertial) ECI-CS (see Fig. 2). While
models intended for distinct phases differ significantly
in general, they have a common structure. The
kinematic part, with x =[ p 0 , v 0 ] 0 , of the state-space
models of a BT has the form
¡¡!
O S P in
While gravity is present in the entire flight of
a BT, it is the dominating, if not sole, force acting
on a coast BT. The following three gravity models,
denoted by a (0)
the (East-North-Up) ENU-CS and p =
G , a (1)
G ,and a (2)
G , are commonly used for
BT tracking.
1) Flat Earth Model : The simplest possible
model of gravity assumes a flat, nonrotating Earth.
In this model, the gravity acting on the target in the
ENU-CS is a constant:
a (0)
·
¸
v
a
_ x =
:
(2)
G =[0,0, ¡g ]
0
(3)
Various models differ from one another in 1) the
choice of models of acceleration a by accounting
for different acceleration components at varying
accuracy levels, and 2) the additional models needed
for describing the evolution of unknowns in the
acceleration components.
We survey dynamic models proposed/used for
the distinct regimes of a BT. The motion phases are
presented in an order according to their sophistication
levels, rather than their chronology in the trajectory.
For each phase, we describe physics (forces or
accelerations) first and then motion models in
different CS.
Compared with [64], this survey is slightly more
tutorial in nature for the benefit of the readers less
familiar with the ballistics or aerodynamics. The
where g is the constant gravitational acceleration of
the Earth.
2) Spherical Earth Model : Assume that the Earth
and the BTs can be represented as point masses at
their centers 1 and that the gravitational forces of the
moon (and stars) can be neglected. Since the target
has a negligible mass relative to the Earth’s mass,
the gravitational acceleration a G is the solution of a
so-called restricted two-body problem, obtained by
Newton’s inverse-square gravity law [7] as
G ( ½ )= ¡ ¹
k½k 2 u ½ = ¡ ¹
k½k 3 ½ (4)
1 This holds if the Earth and the targets are spherically symmetric
with an even distribution of their masses.
LI & JILKOV: SURVEY OF MANEUVERING TARGET TRACKING, PART II
97
AUTHORIZED LICENSED USE LIMITED TO: IEEE XPLORE. DOWNLOADED ON MAY 13,2010 AT 11:53:33 UTC FROM IEEE XPLORE. RESTRICTIONS APPLY.
a (1)
641081518.006.png
where ½ is the vector from the Earth center to the
target, k½k its length, u ½ = ½=k½k its unit vector, and
¹ is the Earth’s gravitational constant.
3) Ellipsoidal Earth Model : More accurate
expressions for the gravitational acceleration can
be obtained by replacing the above spherical Earth
model with an ellipsoidal (or more precisely,
spheroidal) Earth model. Such a more precise
approximation–accounting for the Earth oblateness
by including the second-order gravitational harmonic
term J 2 of the Earth’s gravitational field model–is
[7, 23]
C
induced by the Earth’s rotation [72, 7], which consist
of the Coriolis 2 a (1)
C + a (2)
C and centrifugal a (2)
C terms:
C = −£ ( − £½ )= M ½
(6)
where =[ X , Y , Z ] 0 is the Earth’s angular velocity
vector, v is the target’s linear velocity vector in the
noninertial Earth-fixed CS [75, 7] (Fig. 2), and the
antisymmetric matrix M is given by
2
3
(
0
¡− Z
Y
Μ
G ( ½ )= ¡ ¹
u ½ + 3
r e
k½k
2
4
5
a (2)
k½k 2
2 J 2
M =
Z
0
¡− X
:
(7)
¡− Y
X
0
)
£ [(1 ¡ 5( u 0
½ u Z ) 2 ) u ½ +2( u 0
½ u Z ) u Z ]
For the boost and reentry portions of a BT in a
close vicinity of the Earth and over a short period,
the effect of the Coriolis and centrifugal forces may
be neglected. During the ballistic flight, however, this
effect is usually significant and should be accounted
for, particularly in the case of a long-range BT
(relative to the Earth’s radius).
For simplicity of presentation, let
(5)
where r e is the Earth’s equatorial radius, J 2 is a
correction constant, and u Z is the unit vector along
¡!
O Z I (see Fig. 2). The best-known Jeffery constant
J 2 represents the difference between the polar
and equatorial moment of inertia. It quantifies the
oblateness of the Earth and is approximately equal
to one third of the ellipticity of the Earth. Even more
accurate models are available. For example, the
specifications of the World Geodetic System 1984
(WGS-84) ellipsoid and the Earth Gravitational Model
(EGM 96) are given in great detail in [77].
The boost and reentry phases are relatively
short in range compared with the Earth radius and
take place in a close vicinity of the Earth. Thus
a flat, nonrotating Earth model may be adequate,
particularly in the presence of other more dominating
uncertainties. On the contrary, the coast flight of a
long-range BT comprises a much greater range, and
thus accounting for the Earth sphericity (and even
ellipticity) and rotation is essential. The spherical
model (4) is classical and has been most commonly
used in a variety of BT tracking applications
[90, 51, 55, 85, 29, 99, 56, 5, 9]. It has been used
for BT tracking over a short range and/or period, for
example, as a gravity model as an integral part of
the total acceleration within a boost (boost-to-coast)
motion model [85, 29, 5, 9] or for track initiation
purposes [99]. However, it may be inadequate
for more demanding BT tracking applications, in
particular, precision tracking of coast targets over
a long time period or at a low data rate. For a
long-range coast BT, gravity is the sole or dominating
acceleration and thus needs more accurate models
such as the ellipsoidal Earth or WGS-84 model.
a ( i G , I
g i , noninertial CS , i =0,1,2 (8)
be the total acceleration induced by the Earth, where
a ( i G was given by (3)—(5) and g i = a ( i G
½
a ( i )
E =
expression in an Earth-fixed frame.
¡a C is its
C. Coast Motion Models
In the exo-atmospheric coast phase, no thrust
is applied and no drag is experienced. The motion
may be considered governed by the acceleration a E
induced by the Earth alone, with other factors (e.g.,
perturbations) neglected. Thus, the total acceleration is
a = a E . It follows from (2) that the state-space models
of a coast target with the state vector x =[ p 0 , v 0 ] 0 have
the form
·
v
a ( i )
E
¸
_ x =
(9)
E was given by (8). So, coast modeling of
the BT amounts to selection of an appropriate model
for a E .
1) Inverse-Square Model in ECI-CS :Inths
case, the acceleration part of the state-space model
(9) is described by the inverse-square model (4)
in the E CI-CS as a ( i )
p
E = a (1)
G ( p )= ¡¹p=kpk 3 ,where
X 2 + Y 2 + Z 2 .
A trajectory described by (9) with (4) is confined
to the so-called orbital plane and is a part of a conic
B. Coordinate Accelerations
If the target motion is considered in a noninertial
frame fixed to the Earth (e.g., the ENU-CS, ECF-CS,
and body frames), its relative total acceleration
2 The Coriolis acceleration is induced by the rotation of the Earth
that causes the Coriolis effect–the apparent deflection of a body
in motion with respect to the Earth, as seen by an observer on the
Earth. This effect appears in a noninertial CS [7].
98
IEEE TRANSACTIONS ON AEROSPACE AND ELECTRONIC SYSTEMS VOL. 46, NO. 1 JANUARY 2010
AUTHORIZED LICENSED USE LIMITED TO: IEEE XPLORE. DOWNLOADED ON MAY 13,2010 AT 11:53:33 UTC FROM IEEE XPLORE. RESTRICTIONS APPLY.
includes the coordinate accelerations a C = a (1)
a (1)
C =2 − £v =2 M v , a (2)
where a ( i )
kpk =
641081518.007.png 641081518.001.png
orbit, governed by the Keplerian motion equation [7].
A dynamic model is used in target tracking mainly
for propagating the target state (i.e., state prediction),
known as the Kepler problem in astrodynamics,
and for covariance prediction. Although the model
given by (9) with (4) is highly nonlinear, the state
propagation with it can be done in an efficient
manner, through the algorithm outlined below.
Given x ( t 0 )=[ p 0
G ( p )ofthe
state-space model (9) has the following form [7, 29]:
E = a (2)
2
4
Ã
Ã
Μ
2
!!
3
5
1+ c e
kpk 2
Z
kpk
1 ¡ 5
X
Ã
Ã
Μ
2
!!
G ( p )= ¡ ¹
1+ c e
kpk 2
Z
kpk
a (2)
kpk 3
1 ¡ 5
Y
0 ] 0 at time t 0 the predicted target
state x ( t )=[ p 0 , v 0 ] 0 at t = t 0 + T (with T> 0) is given
by [7]
0 , v 0
Ã
Ã
!!
Μ
2
1+ c e
kpk 2
Z
kpk
Μ
Μ
3 ¡ 5
Z
1 ¡ ¸ 2
kp 0
¸ p
p =
k C
p 0 +
¹ S
v 0
(10)
p
Μ
¹¸ ( ³S¡ 1)
kp 0
1 ¡ ¸ 2 C
kp 0
where c e =(3 = 2) J 2 r e .
This model was chosen in [29], [28] as the coast
model for a 6-state EKF-based ballistic filter in the
ECI-CS implemented in a multiple-model tracking
system. However, as pointed out in [29], a model
mismatch caused by moderate trajectory perturbations
may lead to a track divergence. This effect can be
alleviated by introducing small fictitious zero-mean
process noise. This technique was used in [29] to
adapt the model for the post-boost phase of the
trajectory, which may involve small maneuvers.
In the early work on BT tracking (e.g., satellite
orbit determination [51]), the target dynamics was
usually considered deterministic (i.e., without process
noise). This often led to divergence of the EKF.
Introducing fictitious process noise is an effective
means to account for such factors as model errors,
neglected perturbations, nonlinearities, and computer
roundoff errors. While the state prediction remains
unaffected if additive zero-mean process noise input is
present, the prediction error covariance is increased
by the process noise covariance Q , and thus the
possibility for the EKF to diverge is reduced. The
price paid is a possible accuracy degradation of
the filter when the deterministic model is indeed
adequate. This is closely related to the problem of
noise identification and adaptive filtering. See, e.g.,
[51], [71], [60]. More recent work along these lines
for tracking an orbital target can be found in [44],
where Q was tuned adaptively using the most recent
state estimate and covariance in a gravity-gradient
model with the inverse-square law. Although a
number of ways to choose Q have been proposed
in the literature, in practice Q still remains a design
parameter, which is adjusted/tuned based mostly on
engineering experience and intuition [23, 12]. More
comprehensive discussions of this issue can be found
in [63] and [64].
3) Model with J 2 Correction in ENU-CS :ts
straightforward to obtain the state-space form of the
acceleration model in the ENU-CS associated with
this gravity model as
v =
kkpk p 0 +
k
v 0
¹=kpk and can be computed accurately and
efficiently via the Newton iteration scheme for solving
a so-called time-of-flight equation, as given below
for the case of an elliptical orbit (i.e., a =2 =kp 0
p
kv 0
k 2 =¹> 0) [7]:
1. Initialize
® := 1 ¡ a k p 0
p
k
, ¯ := p 0
0 v 0
¹
¹
¸ := a p
¹ , ° :=
kp 0
p
¹ :
k
2. Repeat the following until jT¡¿j<² :
³ := 2 , C := 1 ¡ cos
p
³
³
p
p
³¡ s in
³
S :=
p
, ¿ := ®¸ 3 S + ¯¸ 2 C + °¸
³
³
d¿
:= ®¸ 2 C + ¯¸ (1 ¡³S )+ °
·
¸
d¿
¡ 1
¸ := +
( T¡¿ ) :
A useful program-like pseudocode (with a few
small typographical errors) of the above algorithm
for all conic (i.e., elliptical, parabolic, and hyperbolic)
trajectories can be found in [99] (see also [89], [19]),
along with the computation of the Jacobian F ( t , t 0 )=
@x ( t ) =@x ( t 0 ) necessary for an extended Kalman filter
(EKF) [51, 4]. Explicit evaluation of the Jacobian by
direct differentiation of (4) can be found in [55]. For a
comprehensive treatment of the theoretical background
and solutions to the Kepler problem, the reader is
referred to [7].
2) Model with J 2 Correction in ECI-CS :A
refined model for the gravitational acceleration is
based on the ellipsoidal Earth model (5). Note that
u p =[ X , Y , Z ] 0 =kpk and u Z =[0,0,1] 0 in the ECI-CS.
a E = g 2 = a (2)
G ( ½ ) ¡ 2 M v¡M ½ (11)
LI & JILKOV: SURVEY OF MANEUVERING TARGET TRACKING, PART II
99
AUTHORIZED LICENSED USE LIMITED TO: IEEE XPLORE. DOWNLOADED ON MAY 13,2010 AT 11:53:33 UTC FROM IEEE XPLORE. RESTRICTIONS APPLY.
Similar to (4), the acceleration part a ( i )
where C , S ,and ³ are known functions of ¸ .The
variab le ¸ is defined through its time derivative by
¸ =
641081518.002.png 641081518.003.png
where ½ = p + r =[ X , Y , Z + krk ] 0 ,with r being the
vector from the Earth center to the origin of the
ENU-CS, and krk = r e + Z S is the sum of the Earth
radius and the altitude of the origin.
Conversion to the radar face CS and the
corresponding RUV-CS (see the Appendix) and of its
application in the EKF can be found in [23]. A similar
model is given in [70] with inverse-square gravity,
an ellipsoidal Earth, and Coriolis and centrifugal
acceleration terms in the radar face CS, along with
the respective EKF with RUV measurements.
4) Inverse-Square Model in ENU-CS :W ththe
spherical Earth’s gravity, the model (11) is simplified
to
jt k ). For the propagation of the
associated error covariance needed in, e.g.,
Kalman filtering, it is proposed in [56] to add, in
effect, component-uncoupled small noise w ( t k )=
[ w X , w Y , w Z ] 0 ( t k )to[ ˆ X , ˆ Y , ˆ Z ] 0 ( t k
jt k ) in the dynamics
equation, where cov( w )=diag( q X , q Y , q Z )isa
design parameter. As a result, what is proposed
is a coordinate-“uncoupled,” state-dependent,
piecewise-constant, non-zero-mean white-noise
acceleration model, using the terminology of [4],
[64]. There seems room for improvement here by a
better way of adding noise or by adding temporally or
spatially correlated noise, but at a price of increased
complexity.
Note that the motions along the X , Y , Z directions
described by this model are not actually uncoupled
over multiple time steps. The coupling arises from the
dependence of the acceleration on the common target
state in (14). This model has been reported in [56] to
provide performance-competitive with that of the EKF
based on the fully coupled nonlinear inverse-square
model-for several BT tracking scenarios with a high
sampling rate (of 4 Hz). It seems worthwhile to try
this model in conjunction with some more precise
acceleration models, rather than (12), as originally
proposed in [56].
The underlying idea of this “acceleration
compensation” [35] approach is not restricted to
coast models or the specific form of the acceleration
model used: The piecewise-constant approximation
canbedirectlyappliedinanysituationwherethe
total acceleration is available as a function of the
position and velocity. Indeed, the same approach
was used in [52] to account for gravity in a simple
kinematic model of gravity turn motion during
boost (see Section IVB2); similarly, in [35] a
thrust acceleration term derived from a (nearly)
constant-acceleration (CA) filter estimates was
included in the total acceleration to predict the
boost motion of a BT; an additional drag-induced
acceleration term was included in [31], along with
the gravity and thrust terms. In general, this approach
is simple for implementation, but its accuracy
may be inadequate in the case of a low sampling
rate.
6) Models in Other Coordinates : Models in
other CS (e.g., radar face CS and spherical CS) can
be obtained by conversion from the ENU-CS. For
example, an explicit expression of the inverse-square
model (neglecting the Coriolis and centrifugal forces)
can be found in [15].
a E = g 1 = a (1 G ( ½ ) ¡ 2 !©v¡! 2 © 2 ½ (12)
where ! is the Earth rotation rate, and
© =
2
4
0 ¡ sin Á cos Á
sin Á 0 0
¡ cos Á 0 0
3
5
(13)
where Á is the latitude of the origin of the ENU-CS.
This follows from = ! [0,cos Á ,sin Á ] 0 and M =
in this case (i.e., X =0).
Like the models considered before, the standard
EKF technique is directly applicable to this model
in the ENU-CS. The linearization needed for error
covariance propagation may not be sufficiently
accurate for relatively long propagation time periods.
For cases with high sampling rates, however, a simple
yet accurate piecewise-constant acceleration model
[56] could be used, as discussed next.
5) Piecewise-Constant Acceleration Model :The
idea of this approach, proposed in [56], is simple
and natural. At each propagation time step t k
jt k )=[ X , ˆ _ X , Y , ˆ _ Y , Z , ˆ _ Z ] 0 ,thatis, a ( t k
jt k )= a ( x ( t k
jt k )).
For example, if (12) is used to model the total
acceleration, then
a ( t k
jt k )= g 1 ( x ( t k
jt k )) :
(14)
jt k )is
constant over the time period [ t k , t k +1 ), the following
discrete-time dynamics model along the X direction is
clearly obtained
·
X ( t k +1 )
_ X ( t k +1 )
¸
·
1 t k +1
¡t k
¸·
X ( t k )
_ X ( t k )
¸
=
0 1
·
¸
¡t k ) 2 = 2
t k +1
ˆ X ( t k
+
jt k ) :
¡t k
III. REENTRY
Likewise for Y and Z directions.
This state-dependent piecewise-constant
acceleration model is simple and straightforward
During reentry, two significant forces are
always present: the gravity and atmospheric drag. If
maneuvers are possible, aerodynamic lift must also
100
IEEE TRANSACTIONS ON AEROSPACE AND ELECTRONIC SYSTEMS VOL. 46, NO. 1 JANUARY 2010
AUTHORIZED LICENSED USE LIMITED TO: IEEE XPLORE. DOWNLOADED ON MAY 13,2010 AT 11:53:33 UTC FROM IEEE XPLORE. RESTRICTIONS APPLY.
for application. It is used for state prediction
to get x ( t k +1
!t k +1 ;
first sample (i.e., compute) the continuous-time
total acceleration process at the estimated point
x ( t k
As such, under the assumption that the target motion
is uncoupled along X , Y ,and Z directions, and the
acceleration a ( t )=[ ¨ X ( t ), ¨ Y ( t ), ¨ Z ( t )] 0 = a ( t k
( t k +1
Zgłoś jeśli naruszono regulamin