Optimizing The Receiver Maneuvers For Bearings-Only Tracking

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Automatica 35 (1999) 591—606

Brief Paper
Optimizing the receiver maneuvers for bearings-only tracking
J.-P. Le Cadre *, S. Laurent-Michel 夽
IRISA/CNRS, Campus de Beaulieu, 35 042, Rennes, France
 Mastere EST, ENST, 46 rue Barrault, 75 634, Paris, cedex 13, France
Received 12 August 1996; revised 27 January 1998; received in final form 29 June 1998

Abstract

This paper deals with the optimization of the observer trajectory for target motion analysis. The observations are made of estimated
bearings. The problem consists in determining the sequence of controls (e.g. the observer headings) which maximizes a cost functional.
This cost functional is generally a functional of the FIM matrix associated with the estimation of the source trajectory parameters.
Further, note that these parameters are only partially observed. The determinant of the Fisher information matrix (FIM) has all the
desirable properties, the monotonicity property excepted. This is a fundamental difference with ‘‘classical’’ optimal control. The
analysis is thus greatly complicated. So, a large part of this paper is centered around a direct analysis of the FIM determinants. Using
them, it is shown that, under the long-range and bounded controls hypotheses, the sequence of controls lies in the general class of
bang-bang ones. These results demonstrate the interest of maneuver diversity. More generally, they provide a general framework for
devising a strategy for optimizing the observer trajectory, only based on the observations (i.e. the bearing rates).  1999 Elsevier
Science Ltd. All rights reserved.

Keywords: Target tracking; Optimal control; Optimization; Bang-bang control

1. Introduction vering sources constitute an important area of present


research (Chang and Tabaczinski, 1984; Blom and Bar-
Passive bearings-only tracking (BOT for the sequel) Shalom, 1988). In the case of a maneuvering source,
techniques are used in a variety of applications (Nardone a leg-by-leg hypothesis for the source trajectory is quite
et al., 1984; Aidala and Hammel, 1983) like sonar, in- common in the sonar context.
frared tracking (Barniv, 1985), optronic or electronic war- In the BOT context, the source trajectory is only
fare. In the passive sonar context, the basic problem partially observed through noisy non-linear measure-
of target motion analysis (TMA) is to estimate the traject- ments (estimated bearings). Here, a moving observer
ory of a moving object (e.g. position and velocity for passively monitors noisy bearings to estimate a source
a rectilinear motion) from noise-corrupted sensor data. trajectory. A great deal of work has yet been devoted to
Classical bearings-only TMA methods are restricted to the analysis of the observability (Nardone and Aidala,
constant motion parameters (usually position and velo- 1981; Le Cadre and Jauffret, 1997). However, this is
city) (Nardone et al., 1984) even if extensions to maneu- a binary (yes/no) analysis and a practical fundamental
question remains: if the system is observable what is the
—————
accuracy of the source trajectory estimate? The TMA
*Corresponding author: Tel.: #33 2 99 84 72 24; fax: #33 2 99 methods are now well understood and their performance
84 71 71; Elec mail: lecadre@irisa.fr. has been carefully investigated, at least for simple scen-
 This paper was not presented at any IFAC meeting. This paper was arios (Nardone et al., 1984; Trémois and Le Cadre, 1996).
recommended for publication in revised form by Associate Editor It is quite evident that the performance of any TMA
HenkNijmeijer under the direction of Editor Tamer Bas,ar. This work
has been supported by DCN/Ingénierie/Sud. (Dir. Const. Navales),
France —————

Present address: Synopsis, 700 East Middlefield Road, Mountain  A leg is a portion of the trajectory where the velocity vector is
View, CA 94 043— 4033, USA constant

0005-1098/99/$—see front matter  1999 Elsevier Science Ltd. All rights reserved
PII: S 0 0 0 5 - 1 0 9 8 ( 9 8 ) 0 0 1 8 3 - 6
592 J.-P. Le Cadre, S. Laurent-Michel/Automatica 35 (1999) 591— 606

method is highly depending on the observer maneuvers gence, possibly to a local extremum. However, a system
Mc Cabe, 1985; Fawcet, 1988). In the past, the import- of 18 nonlinear coupled differential equations is needed
ance of this problem has been recognized even if efforts which leads to a certain numerical burden. The approach
were, for a large extent, devoted to the considerably of Passerieux and Van Cappel (1991, 1998) is also quite
simpler case of bearings-only localization (i.e. the source brilliant but seems less demanding. The aim of such
is fixed). For instance, Liu (1988) and then Hammel et al. approaches is mainly to provide a catalogue of recom-
(1989) have considered a lower bound of the determinant mended maneuvers (see Section 4).
of the two-dimensional Fisher information matrix (FIM), In fact, the optimal control approaches are very at-
allowing to approximate the problem within the general tractive. Unfortunately, as we shall see later, the det
framework of optimal control (see Section 3). functional does not have the monotonicity property (see
The general problem consists in optimizing controls, Sections 2 and 4) so it is not at all true that adding
which are kinematic parameters of the observer, in order a control optimal for the time t to a control sequence
to optimize the statistical behavior of the TMA. A classi- optimal up to time t#1 will yield an optimal control
cal approach consists then in considering the (FIM) and sequence up to time t. This explains, for a large part, the
more precisely its determinant. The choice of the determi- complexity of the analysis. Practically, this means that
nant functional is reasonable (Helferty and Mudgett, our problem cannot be treated (in its full generality) with
1993; Jauffret and Musso, 1991). This is a common cost the methods of optimal control. So, a large part of the
functional in the estimation literature. The relations be- paper is centered a direct analysis of the FIM determi-
tween the criteria derived from various functionals have nants which will play the central role. More precisely, we
been thoroughly studied in the optimal design literature. shall show that using elementary multilinear algebra,
Furthermore, we shall show that, under hypotheses rea- accurate approximations of det(FIM) may be obtained.
sonable in the BOT context, the maximum of det(FIM) is More specifically, we shall prove that det(FIM) may be
attained when the sphericity criterion is maximum. More approximated by a functional involving only the success-
generally, note that this problem is, of course, not specific ive source bearing rates, thus yielding the general form of
of TMA and arises in a great variety of applications, the optimal controls (observer maneuvers). In particular,
mixing estimation and control. A good example is the it will be shown that, under the long-range and bounded
optimisation of tracking performance (Kershaw and controls hypotheses, the sequence of optimal controls
Evans, 1994). The absence of any a priori knowledge is a bang-bang one. These results demonstrate the
about the source trajectory is also a great source of interest of maneuver diversity. More generally, they
difficulty. This is solved by considering the problem in its provide a general framework for optimizing the observer
natural framework, i.e. that of modified polar coordi- trajectory and constitute the major result of this
nates, thus separating the observable and unobservable paper.
parameters as well as their respective effects. First, a general presentation of the problem will be
Another reasonable assumption consists in modelling given in Section 2. Connections with optimal design
the observer trajectory by a sequence of legs. A leg is approaches and optimal control are also included. The
a part of the trajectory where the motion is rectilinear applications of optimal control theory are considered in
and uniform (constant velocity vector) (Nardone et al., Section 3. Section 4 constitutes the core of the paper. It
1984). Then, the problem consists in determining the relies upon a direct analysis of the FIM determinant.
sequence of observer headings which optimize a cost First, a constant source bearing rate is considered. Then,
functional (Holtsberg, 1992). For instance, Fawcett using the same formalism, these results are extended to
(1988) has considered a two-leg observer trajectory and the case of time-varying (piecewise constant) source bear-
has determined the heading of the second leg which ing rates. The optimisation problems are presented in
maximizes the accuracy of the source range estimation. detail. This is followed by a geometric interpretation and
Jauffret and Musso (1991) and Hammel et al. (1989) have general conclusions.
extended this work to an arbitrary number of observer Standard notations will be used throughout this
legs (up to 20 for Hammel, 1988). Even if this approach is paper:
mainly computational, interesting and thorough insights
have been obtained by this way. Another view of the E a bold letter denotes a vector while a capital standard
problem has been provided by Olsder (1984) and Pas- letter denotes a matrix,
serieux and Van Cappel (1991, 1998). In their ap- E a capital calligraphic letter generally denotes a sub-
proaches, the FIM components are included in the space or a subset of vectors,
source state. The corresponding methods lie in the class E the symbol (*) means transposition,
of optimum design approaches, thus assuming that the E r and r represent the relative x and y coordinates,
V W
source trajectory is known. In Olsder’s work (1984), the v and v denote the relative x and y velocities,
V W
idea is that, given a certain maneuver, a better maneuver E t or k is the time variable, r is the distance (range), h is
is found and this procedure is repeated up to conver- a bearing and u a heading,
J.-P. Le Cadre, S. Laurent-Michel/Automatica 35 (1999) 591— 606 593

E the symbol " G means approximation at the ith order, In the above formula t is the time at the kth sample,
I
E det represents the determinant, tr stands for the trace, while the vector u "(0, 0, u (k), u (k))* accounts for the
I V W
diag denotes a diagonal matrix. effects of observer accelerations (or controls). We denote
E Id is the n-dimensional identity matrix, the symbol cov U _(u , 2 , u ), the sequence of observer controls up to
L L  L
stands for the covariance matrix, time n. Eq. (2.2) assumes that between t and t the
I I>
E BOT: bearings-only tracking; RUN: rectilinear uni- source motion is rectilinear and uniform, this hypothesis
form motion; TMA: target motion analysis; FIM: will be taken along the whole paper. All along the paper,
Fisher information matrix; PMP: Pontryagin max- the vector X denotes the relative state vector at a given
imum principle. instant. An usual choice for the reference time is 0, in this
case X is X . For the sake of brevity, the reference time

will be generally omitted. Indeed, it is a fundamental
2. Problem formulation result that our analysis is independent of the reference
time (cf Proposition 2).
The general notations are identical to those of the As usual in passive TMA (Nardone et al., 1984), the
reference paper (Nardone et al., 1984). The physical para- available measurements are the estimated angles hª (bear-
I
meters are depicted in Fig. 1. The source, located at the ings) from the observer to the source, so that the observa-
coordinates (r , r ) moves with a constant velocity vec- tion equation stands as follows (w is the measurement
V W I
tor v (v , v ) and is thus defined to have the state vector: noise):
V W
X _[r , r , v , v ]*. (2.1) hª "h #w , (2.3)
 V W V W I I I
The observer state is similarly defined as with

 
r (k)
X _[r , r , v , v ]*, h "tan\ V .
 V W V W I r (k)
W
so that, in terms of the relative state vector X defined by
The measurement noise w is usually modelled by an
I
X"X !X _[r , r , v , v ]*, i.i.d. zero-mean, gaussian process with a given variance
  V W V W p. The four-dimensional state equation (2.2) and the
the discrete-time equation of the system (i.e. the equa- non-linear measurement equation (2.3) define the bear-
tion of the relative motion) takes the following form ings-only motion analysis process (TMA). Given the his-
tory of measured bearings #ª _+hª ,L the likelihood
(a_t !t "cst):
I> I G G
function is (Nardone et al., 1984)
X "FX #u , (2.2)
I> I I P(#ª "X)"cst exp [! ##ª !#(X)# ] ,
 
where
#(X) defined by Eqs. (2.2) and (2.3),

   
Id aId 1 0 ##ª !#(X)# _ (#ª !#(X))*&\(#ª !#(X)), (2.4)
F"'(k, k#1)"   , Id _ . 
0 Id  0 1 &"diag (p).

The maximum likelihood estimate (MLE) Xª is the
solution to the likelihood equation

Xª "arg maxX log [P (#ª "X)]. (2.5)

The above equation does not have explicit solution.


So, the following Gauss—Newton algorithm is usually
considered (i is the iteration index):

  
*#(X) * *#(X) \
Xª "Xª !o &\
G> G G *X *X X"Xª
G

 
*#(X) *
; &\ (#ª !#(Xª )) (2.6)
*X G

with o being the step size of the algorithm.


G
The calculation of the gradient vector of the compo-
Fig. 1. Typical TMA scenario. nents of #(X) is easily derived from Eqs. (2.2) and
594 J.-P. Le Cadre, S. Laurent-Michel/Automatica 35 (1999) 591— 606

(2.3), yielding Denoting G  the gradient vector of the log-likelihood


I
functional (2.4), i.e. .
r (k) r (0)#kv # u
tan(h )" V " V V J J V ,
I r (k) r (0)#kv # u 1
W W W J J W G" (cos (h ),!sin (h ), k cos(h ),!k sin (h ))*,
I p r I I I I
and consequently I I
the problem is to determine the sequence of controls
* 1 *h 1
tan(h )" I " , +u , 2 , u , (denoted U ) such that
*r (0) I cos(h ) *r (0) r (k)  L L
V I V W
  
L
so that U Parg max det G G* , (2.11)
L I I
I
*h cos h r (k) cos h cos h
I & I" W I" I. (2.7) r (0)#kv # I u
*r (0) r (k) r r (k) r tan(h )" V V J J V .
V W I W I I r (0)#kv # I u
W W J J W
The other components of the partial derivatives are ob-
In Eq. (2.11), the gradient vector G denotes, in fact, the
tained by the same way, giving I
vector GX (X "f (X, U ), see Eq. (2.2)). The difficulty
*h sin h I I I I
I "! I, and the originality of the above problem stem from the
*r (0) r two following facts. First, the source motion is unknown
W I
which means that the reference state vector X is un-
*h k k
I" cos h " cos h , known. The problem is to optimize its estimation. So,
*v r (k) I r I (2.8)
V W I practically, the sequence of bearings is unpredictable.
*h k Second, we deal with a global optimization problem
I"! sin h . which means that we seek for an optimal sequence of
*v r I
W I controls maximizing a cost function based on the whole
Collecting the previous results, the partial derivative FIM matrix. Of course, the problem is drastically eased
matrix of the bearing vector #(X) with respect to the by considering an additive (matrix) cost functional, but
state components is deduced (Nardone et al., 1984), (as we shall see later) if the problem becomes far simpler
yielding the associated solutions optimized observer trajectories)

 
cos h sin h cos h sin h perform quite poorly. Moreover, in this context, our
 !   !  problem can be treated by means of the dynamic pro-
r r r r
*#(X)     gramming principle. Unfortunately, it necessarily re-
" $ , (2.9)
*X quires that the cost functional f (from H , the vector
cos h sin h n cos h n sin h L
L ! L  ! L space of n-dimensional Hermitian matrix to 1) satisfies
r r r r
L L L L the following motonicity property, denoted MDP
(Matrix Dynamic Programming Property) and defined
where +h ,L represent the source bearing at the instant
G G below:
i and +r , the source-observer distance. In Eq. (2.9) the
G
reference time is the instant 0 (see Proposition 2). Con- Definition 1. The function f (H P1, differentiable, C)
sider now the case of a non-maneuvering source (con- L
has the MDP if the following implication holds, whatever
stant velocity vector), then the calculation of the FIM is C3H :
a routine exercise yielding, under the Gaussian assump- L
tion (Nardone et al., 1984) f (B)'f (A) N f (B#C)'f (A#C).

   
*#(X) * *#(X)
FIM (X, U)" &\ . (2.10) An interpretation of this definition in terms of dynamic
*X *X programming (maximization) is the following type of
We shall use this formulation of the FIM along the inequality (Le Cadre and Trémois, 1997):

 
whole paper. Actually, the performance of any TMA I
algorithm is dramatically related to the observer maneu- maxUM * f GX G*X 4max [ f +GX G*X #F (X , UM * ,]
I G G SI I I  I>
vers. So, optimizing these maneuvers represents the main GL
problem in TMA. It is therefore not surprising that which must be valid for the strategy UM * , optimal up to
I
a great deal of work has been devoted to this subject. The time k, and for k"n!1, 2 , 0. Roughly, the MDP
problem then consists in determining a sequence of con-
trols maximizing a cost function. Since we deal with the
estimation of the source trajectory (the vector X, in fact) —————
 *#(X)/*X"(G , 2 , G )*.
it seems reasonable to consider that the cost function  L
 X is X if the reference time is 0.
is a functional of the FIM, leading to the following 
 F denotes a FIM matrix, UM the optimal sequence of controls from
I
problem: n to k.
J.-P. Le Cadre, S. Laurent-Michel/Automatica 35 (1999) 591— 606 595

appears as a material form of a ‘‘comparison’’ principle. stating that the three following characterizations of a
So, it plays a fundamental role in dynamic programming. D-optimal design m* are equivalent:
A fundamental question consists in determining the func-
(i) m* maximizes det (M(m)),
tionals f having the MDP property. An answer is pro-
(ii) m* minimizes dM (m) _ supX3s d(X, m), where
vided with the following result.
d(X, m)"G*
X M(m)\GX , (2.14)
Proposition 1. ¸et f satisfy the MDP property, then
(iii) dM (m*)"dim(X).
f (A)"g (tr(AR)).
These equivalences were first proved by Kiefer using
where g is any monotonic increasing function and R is game-theoretic methods. A general version of this the-
a fixed matrix. orem (with a simpler proof ) was later given by Whittle
(1973). Moreover, a characterization of the D-optimality
We refer to Le Cadre and Trémois (1997) for the proof in terms of invariance of ranking has been obtained. All
of Proposition 1. It is simply based on the fact that if these considerations plead for considering D-optimal

f (A) and
f (B) are not colinear, then their respective design despite the fundamental difficulties mentioned
orthogonal subspaces are distinct which implies that above. Various algorithms for the calculation of D-opti-
there exists a matrix C for which the MDP is not satis- mal design have been proposed (Atwood, 1973), a general
fied. Therefore,
f (A) and
f (B) must be colinear, what- formulation stands as follows (i is the iteration index,
ever A and B. This is a very strong property. In turn, this dX the measure concentrated on X and the ' design
yields the general form of f. Consider for instance functional):
f (A)"log det A, then (Lancaster and Tismenetsky) for
im(i)#dX(i)
a non-singular matrix A, m(i#1)" ,
i#1
Df (C)"tr (A\C)"(
*f (A), C), where

and we see immediately that f does not have the MDP X(i)"arg max D(X, m ),
property. The same remark is valid for functionals as G
simple as f (A)"tr(A\). and
Actually, our problem presents strong similarities with D(X, m)"'(m, dX), (2.15)
the theory of optimal experiment design (Whittle, 1973).
More precisely, this general problem consists in deter- '(m, g)"limeP0 e\ ['+(1!e)m#eg,!'(m)].
mining the values of the experiments X , or more gener- Let us now consider the trace of the FIM; direct calcu-
I
ally a continuous distribution m(dX) in a design space lations yield
which optimize the estimation of an unknown parameter
conditioning the observation (X , here). So, the following 1#k
 tr (G G*)"#G #" ,
matrix M(m) is considered: I I I p r
I I
L whence
M(m)+(1/n) GX G*X ,
I I
I , (1#k)
tr (FIM)" . (2.16)
pr

" GX G*X m(dX). (2.12)  I
Thus, since tr(FIM) is independent of the bearings
The design problem is to determine m so as to maxi- sequence +#,, functionals involving the trace functional
mize some functional ' of M(m). Various choices for are quite simple. However, the associated optimization
' have been proposed in the literature, the two common- problems are restricted to minimizing a functional of the
est ones being those of C-optimality and D-optimality, source-observer distance. Fortunately, for our problem
for which, effectively: (see Eq. (2.11)), the above result means that maximizing
det(FIM) reverts to optimize (as a by-product) the FIM
'(m)"!C*M(m)\C, conditioning.
(2.13) In this spirit, an elegant approach (Olsder, 1984;
'(m)"log (det(M(m)).
Passerieux and Van Cappel, 1991) for this particular
Now, for the case of D-optimality in particular, consider- optimal design problem consists in considering it as an
able use has been made of the equivalence theorem, optimal control problem where the cost is reduced to
Q"det [F(¹)]. Then, the state vector comprises the
FIM elements. However, we note that, basically, the cost
————— functional does not have an integral part. So, it is easily
 C is a given vector. shown that the Pontryagin Maximum Principle (PMP
596 J.-P. Le Cadre, S. Laurent-Michel/Automatica 35 (1999) 591— 606

for the rest) simply reverts to a direct maximization of det Invoking the Minkowski’s inequality, which is valid
(FIM) as described below. for n-dimensional positive-definite matrices A and B, i.e.
Actually, a simple and intuitive approach, to this prob- [det (A#B)]L5(det A)L#(det B)L, we obtain
lem, is to model the observer trajectory by a certain
[det (F #M )]5(det F )#(det M )
number of legs and to optimize the various headings I\ I I\ I
(+u ,) of the observer (Hammel, 1988). Rather surpris- so that
I
ingly, it may be easily shown that this intuitive approach
1 I
and the optimal control based method (Passerieux and [det (F )]5(det F )# [det (M )]. (3.2)
van Cappel, 1991) (discrete time formulation) leads to an I>  2 H
H
identical optimization problem. However, we stress that It now remains to calculate det (M ). Elementary calcu-
this requires the knowledge of the gradient vectors +G , H
I lations yield
and thus implicitly assumes that the source trajectory is
known. Furthermore, local extrema are quite likely. So, det (M )"det [(G , G ) (G , G )*]
H H H> H H>
the object of this approach is mainly reduced to provide "det [Gram (G , G )]
a databank of recommended maneuvers. H H>
In order to remedy for the previous problems, we shall 1
" sin (h !h ). (3.3)
first consider integral approximations of the FIM deter- pr r H H>
H H>
minant. It will then be possible to use efficiently the PMP
A lower bound of the FIM determinant has thus been
formalism (see Section 3). However, the interest of such
derived. An attractive approach consists then in consid-
approximations is relatively limited, essentially because
ering the maximization of this lower bound instead of the
the integral cost, derived in this way, are very poor
direct maximization of det (FIM). The above calculations
approximations of the exact determinant. This is parti-
(see Eq. (3.2) and (3.3)) suggest to consider the following
cularly true when the observer is maneuvering. So, a di-
integral cost (Liu et al., 1988):
rect analysis of the FIM determinant is required (see
2 hQ (t)

Section 4). This constitutes the conerstone of this paper.
C " dt. (3.4)
2 2pr(t)

3. Optimal control theory approaches Using the transversality and stationarity conditions
(for the Hamiltonian), the costates are first determined,
This section is divided in two parts. The first one deals, leading to consider a rather intricated system of differen-
in the localization context, with the maximization of tial equations involving rR , r̈, hQ , h® and u. Rather sur-
a lower bound of the FIM determinant. The basic tool is prisingly however, assuming v (the observer velocity
the Minkowski’s inequality which is again used, in the modulus) constant, an explicit resolution of the optimal-
second part, to derive a lower bound for the moving ity conditions can be obtained, simply yielding (Hammel
source case. et al., 1989) (u is the optimal control)
uR "!2hQ . (3.5)
3.1. Optimal observer trajectory for localization *
This results in a feasible control since an estimation of
In fact, an important part of previous works has been hQ can be obtained from the estimated bearings. The value
devoted to the localization problem, i.e. the restriction of of the Hamiltonian along an optimal trajectory is con-
the TMA problem to the fixed source case. First, we stress stant (Hocking, 1997) and stands as follows:
that this case is considerably simpler in comparison with v
the moving source one. However, the interesting feature H "! , (3.6)
 rhQ
of this approach is that an integral approximation of
the cost functional has been considered. Let us present which means that the product rhQ is constant (v constant).
briefly this approximation. An optimized trajectory thus represents a trade-off be-
At first, the analysis is restricted to the estimation of tween the range and the bearing rate. So, we can consider
the source position (r , r ). The FIM matrix is thus two- that the observer heading is, at first, approximately equal
V W
dimensional. Let us consider the temporal evolution of to the bearing. Then, as the source-observer range de-
the FIM, i.e. creases, the bearing rate increases. The observer is thus
‘‘spiralling’’ around the source.
F "F #M and M "G G*#G G* ,
I> I\ I I I I I> I>
—————
where  The symbol ‘‘Gram’’ denotes a Grammian matrix (Lancaster and

 
1 cos h
G" I . (3.1) Tismenetsky, 1985)
I pr !sin h  (sin (h !h )&hQ  ).
H H> H
I I
J.-P. Le Cadre, S. Laurent-Michel/Automatica 35 (1999) 591— 606 597

3.2. Optimizing the observer trajectory for TMA assumed to be constant. Further, we consider that the
diagonal noise matrix & is proportional to the identity
Obviously, this lower bound approach may be ex- (i.e. &"pId). Even if the first hypothesis seems rather
tended to a moving source. The only changes ( for the cost restrictive, we shall see later (see Section 4.4) that the
functional) and the FIM structure and dimension which effects of range and bearing-rate variations are de-
result in (see Section 4): coupled, allowing us to analyze separately their effects.
Furthermore, the effects of range variations are concen-

 
sin (hQ ) 
det (M )+16 H , trated in a multiplicative term, factor of the determinant.
H rp A direct analysis of the effect of the observer controls,
H H
so that, C becomes based on the discrete-time equation of motion (Eqs.
2 (2.2)—(2.12)) seems unfeasible. We shall thus consider
2 hQ (t)

a simplified model of the source motion:
C " dt. (3.7)
2 2pr(t)
 h "h #jhQ # u , (4.1)
G>H G I
We note that the only change in Eq. (3.7) (versus Eq. (3.4)) I
is the exponent of hQ . Again, this problem can be investi- where hQ is the bearing rate (for a given reference time),
gated by the optimal control approach. However, a main and u is the bearing rate change corresponding to an
difficulty appears. The source is moving, so we cannot observer maneuver (control). For the sequel, the controls
assume that the modulus of the (relative) observer velo- will be the observer bearing-rate changes u .
I
city is constant. The analysis of optimality conditions is The aim of this modelling is to obtain an explicit
considerably complicated by the presence of additional analysis of the FIM determinant. Obviously, the effects of
terms (related to source motions) in the expressions of the observer maneuver are only indirectly analyzed.
bearing and range-rates, thus requiring the use of numer- However, the bearing rates are directly related to the
ical methods (Teo et al., 1991). In order to simplify the cartesian parameters (see Appendix A). Overall, the bear-
calculations, a reasonable assumption is that the ob- ing rates being estimable from the sensor outputs, we can
server velocity is far greater than the source one. thus derive feasible methods for the observer trajectory
Then, the system Hamiltonian is optimization. ¹he fundamental interest of this approach
lies in the fact that no a priori knowledge of the source


hQ  trajectory is assumed. More generally also, the problem
H"! #j v sin u#j v cos u. (3.8)
r   is only partially observable. For the Modified Polar
Coordinate coordinates (Aidala and Hammel, 1983),
The costates j and j are explicited by the method
  only +h, hQ , rR /r, are available measurements. Further,
presented in Appendix B, the optimal control u is then
* we shall prove that det(FIM) is independent of h
deduced:
(cf. Proposition 2). Then, since the effects of bearing-rate


rR  changes and range variations are decoupled, it is quite
4 #uR hQ "!1. (3.9) natural to concentrate our efforts on the effects of bearing
r *
rate changes.
The (constant) value of the system Hamiltonian along We shall denote by F the FIM corresponding to an
I 
an optimal trajectory is arbitrary reference time k and four consecutive measure-

ments, h , 2 , h . Then the FIM F takes the fol-

 
1 rR  I I> I 
H "! hQ #2 . (3.10) lowing form:
 r r
F "(pr)\G G* ,
An optimized observer trajectory is thus a trade-off be- I  I  I 
tween the three terms r, rR /r, hQ . The term (hQ #2 (rR /r)) where
decreases as the square of the source-observer distance. G "(G , G ,G ,G ).
Numerous integral criteria may be considered but not I  I I> I> I>
any seems really suitable for TMA. Actually, as it will be and G is the gradient vector of the observation h w.r.t.
I I
shown in the next section, it is impossible to approximate X , i.e.

conveniently a relevant cost functional (e.g. det F or G "(cos h ,!sinh , k cos h ,!k sin h )*. (4.2)
tr (F\)) by an integral one. I I I I I
Assuming G invertible, we have
I 
det(F )"(pr)\(det G ).
4. A direct analysis of the FIM determinant I  I 
Of course, our attention is not limited to four measure-
For the sake of simplicity, the following assumptions ments per leg. So, the previous calculations will now be
are made along this section. First, the distance will be extended to any number of measurements. Let l be the
598 J.-P. Le Cadre, S. Laurent-Michel/Automatica 35 (1999) 591— 606

number of consecutive measurements and consider now In the same spirit, the vector G may be written as
I
the (4;4) FIM F l (l54) defined as in Eq. (4.2) by
I G "¹I E,
I 
F l"(pr)\G l G* l ,
I I I where

   
where ¹ 0 cos h #sin h
¹_  and ¹ _ ,
 ¹ ¹  !sin h cos h
G l"(G , G , ,G ), l50. (4.3)  
I I I> 2 I>l
h _h /k , E"(1, 0, 0, 0)* (4.7)
Using classical properties of multilinear algebra, I 
namely the Cauchy—Binet formula (Lancaster and Now the following properties are instrumental:
Tismenetsky, 1985), det (F l) is given by the following
I E the matrices R and ¹ are rotation matrices, hence
 
formula: they commute,
det (F l)"(pr)\ [det (G )], E det (R )"det (R )"1.
I #  
# E det (¹ )"det (¹ )"1.
where E is the index subset defined by  
The matrices R and ¹ then also commute and using
 
E"+i , i , i , i , s.t. 14i (i (i (i 4l, this property det G becomes
        #
and det G "det (¹I RG E, ¹IRGE, ¹IRGE, ¹IRGE)
#        
G "(C , C , C , C ), C _G "det (¹I) det (RGE, RGE, RGE, RGE)
# G G G G GH I>GH
. (4.4)     
"det (E, RG\G E, RG\G E, RG\G E). (4.8)
We stress that the above formula plays a central role in   
the analysis of the FIM determinant. Furthermore, the following property has thus been
proved in passing: det G is independent of k and h .
#  I
4.1. The case of constant bearing rate This remarkable property is due to the basic properties
of the determinant and the structures of the matrices
In Eq. (4.4) C stands for the i th column of the matrix R and ¹ . )
GH H  
G. Considering, for instance, a first-order expansion of
the bearings h (i.e. h  h #ihQ ), the calculation of The above determinant itself (i.e. det G "
I>G I>G " I #
det(F l) is reduced to the calculation of the determi- det (E, RG E, RH , E, RI E)) can now be easily calculated by
I   
nants det (G ). Now each of these determinants is the means of exterior algebra (Yokonuma, 1992; Darling,
#
determinant of a 4;4 matrix. Its calculation is greatly 1994), yielding the following simple and general result:
eased by using the following basic result.
Proposition 3. ¹he following equality holds:
Proposition 2. ¸et E be the first vector of the canonical
det G "j(k!i) sin ((k!j)x) sin (ix)#i(k!j)
basis of 1, (E"(1, 0, 0, 0)*), then the following equality #
holds: sin ( jx) sin ((i!k)x),
where
det G "det (RG E, RG E, RG E, RG E). (4.5)
#     i"i !i , j"i !i , k"i !i . (4.9)
     
Proof. Consider the determinant detG (see Eq. (4.4))
#
where as previously, E"+i , i , i , i , and i (i ( Proof. The calculation of det G is greatly eased by using
#
      exterior algebra. First, let us briefly recall the definition
i (i , then
  of the exterior powers of a vector space. Let » be an n-
det G "det (G , 2 , G )
# G G dimensional vector space over 1, then "» consists of all
formal sums a (U V ), where the ‘‘wedge product’’
"det (RG G , RG G , RG G , RG G ),
  G G G H
 I  I  I  I UV is bilinear and alternate. This definition is straight-
where forwardly extended to higher exterior powers (Darling,
1994). For any basis +V , 2 , V , of », the set of p-vectors
 L
   
R 0 cos hQ sin hQ +V 2V , i (2(i 4n, forms a basis of the
R _  and R _ (4.6) G GN  N
 R R  !sin hQ cos hQ n!/(n!p)!p!-dimensional vector space "N». In particular,
 
"1 is one-dimensional, and throughout the paper we
make intensive use of the isomorphism "1,
————— "1"1. The exterior algebra formalism thus ap-
 Note that in Eq. (4.3) the source-observer distance is again assumed pears as an economical way to conduct determinant
to be constant. calculations.
J.-P. Le Cadre, S. Laurent-Michel/Automatica 35 (1999) 591— 606 599

The canonical basis of 1 is denoted +E , 2 , E ,. For order expansions of h is quite straightforward but not
  I>G
the coherence of notations, the vector E (Eqs. (4.4)—(4.8)) truly enlightening.
is identified with E . Then, the components (denoted

a , b , c ) of the exterior products E RG E , in the Remarks. (1) If a third-order expansion of h is con-
      I>G
‘‘reduced’’ basis +E E , E E , E E , of "(1) are sidered instead of the first-order one, then the value of
     
straightforwardly calculated and given below: det (FIM) is exactly zero.. This corroborates the fact that
a "!sin (ix)QE E , the TMA problem is not observable when the observer
   does not maneuver.
b "i cos (ix)QE E , (4.10) (2) However, the BOT problem is observable if multiple
  
c "!i sin (ix)QE E . measurements are available (at each time). In this case,
   bounds derived from Eq. (4.13) are accurate.
Similarly, the components of RH E RI E , in the (3) In fact, a small variation model for the bearing rate,
   
‘‘reduced’’ basis +E E , E E , E E , are i.e. hQ "hQ #g (g w.g.n. N(0, q)) yields a value of
      I> I I I
a "jk sin (( j!k)x)QE E , the type (4.13), Eq. (4.13) roughly appears as an
   upper bound of det (FIM). More precisely, we obtain
b "(k!j) sin ( jx) sin (kx)QE E , (4.11) $ [det (FIM )]+(16c /r) exp (! q) (sin (hQ )).
   R R>  
c "j sin ((k!j)x)!(k!j) sin ( jx)QE E .
   It has thus been shown that det (F ) is proportional to
I J
The determinant det G is deduced from the above
# l(sin h/pr). As practically, h is very small (see Appendix
calculations, by considering the sum of the coefficients A), this means that det (F ) remains very small as far as
of the vector E E E E which spans the one- I J
    no real observer maneuver occurs. So, we shall now inves-
dimensional space "(1), i.e. tigate the effects of a bearing-rate change.
det G "a a !b b #c c ,
#       4.2. The case of bearing-rate change
"j(k!i) sin ((k!j)x) sin (ix)
#i(k!j) sin ( jx) sin ((i!k)x). ) (4.12) We shall now quantify the effects of observer maneu-
vers. First, the following property is an extension of the
Using Proposition 3 and the Cauchy—Binet formula, previous one to this case.
a general formulation of det (FIM) stands as follows: Consider that the temporal evolutions of the source
bearings on two successive legs are described by the fol-
det (FIM)" 14i(j(l4l [ j(k!i) sin ((k!j)x) sin (ix) lowing two linear models (k _k #j; observer
 
#i(k!j) sin ( jx) sin ((i!k)x)]. maneuvering instant):
h "  h #ihQ on the 1st obs. leg 04i4j,
Practically, the following approximations are easily I>G I  (4.15)
deduced from the above property. h "  h  #mhQ on the 2nd obs. leg m50.
I>K I 
Result 1. The following approximation of det G holds Then the following property holds and extends the
# previous result (Proposition 3):
(x"hQ 1):
(ijk) Proposition 4. Consider the case of two consecutive
det G " (k!i) (k!j) ( j!i) x,
# 3 bearing rates x and y; then we have ( j"i!j)
and therefore det (RG E , RH E , RH RI E , RH RJ E )
         

  
(ijk)  hQ  "(k)( j!l) sin ( jx)sin ((k!l)y),
det (F )+ (k!i) (k!j ) ( j!i) .
I J 14i(j(k4l 3 pr #j(l!k) sin (ky) sin ( jx!ly). (4.16)
(4.13) Furthermore, the following approximation holds:
From Result 1, the following approximation is deduced: *
det (RG E , RH E , RH RI E , RH RJ E )+kl ( l!k) (j!i)y.
det(F )+a\[l(1#l)(l#2l!8) (l#2l!3) *y          
I J


hQ  Proof. In Eq. (4.16), the matrices R and R are
;(2#l)] ,  
pr the bearing-rate matrices (cf. Eq. (4.6) associated
with the bearing-rates x and y. Mimicking the proof of

hQ 
Jl . (4.14)
pr
—————
Using the previous formalism, an extension to higher-  The symbol $ denotes the math. expectation.
600 J.-P. Le Cadre, S. Laurent-Michel/Automatica 35 (1999) 591— 606

Fig. 2. Evolution of the FIM determinant det Fl l .

Proposition 3, we calculate the components (a , b , c ) with


  
of E RHY E in the ‘‘reduced’’ basis +E E , E E ,
       K'0, x_hQ , y_hQ .
E E , of "(1) as well as that of RI E RJ E in  
      This result is quite fundamental for TMA (see Prop-
+E E , E E , E E , (denoted (a , b , c ), yielding
         osition 5) and will be clarified by a geometric interpreta-
a "!sin ( jx), a "kl sin ((k!l )y), tion (see Section 4.3).
 
b "j cos ( jx), b "(l!k) sin (ky) sin (ly), The effects of the observer maneuvers on the FIM
  determinant is now illustrated. Thus, in Fig. 2, det (F l ) is
c "!j sin ( jx), c "k sin ((l!k)y)!(l!k) sin (ky) cos (ly). J 
  plotted versus the measurement index. The scenario para-
The first part of Proposition 4 is obtained by calculating meters are described in the Fig. 2. The observer maneuver
the scalar a a !b b #c c . Expliciting the scalar induces a dramatic increase of the FIM determinant
      which is precisely located at the maneuver instant.
*a *b *c
a !b #c  , For the two bearing-rate case, the expansion of
 *y  *y  *y det F l , l  is
I  

 
yields the second part. 1 
F l ,l K P (l , l )y\GxG\ , (4.19)
From Eq. (4.16), the following approximations are I   (pr) G  
easily deduced: G
where the polynomials +P (l , l ), are detailed in
det (RG E , RH E , RH RI E , RH RJ E ) G   G
          Appendix C. From Eqs. (4.18) and (4.19), we note that the
maximum value of det F l ,l is proportional to lhQ 
+kl(l!k) ( j!i)y(x!y#cxy). (4.17) I  
(l Kl , hQ K!hQ ). The result must be compared with
   
The above property allows us to approximate Eq. (4.14). In fact, denoting by Fl (x) the FIM associated
det (FIM) in the case of a maneuvering observer and thus with a constant bearing-rate x and Fl/2, l/2 (x,!x) the
to investigate the effects of the observer maneuvers. In FIM associated with a two-leg observer trajectory (leg 1:
particular, the role of the bearing-rate changes then l/2 meas., bear. rate x; leg 2: l/2 meas., bear. rate !x),
clearly appears. Indeed, since the parameters hQ and Eqs. (4.13) and (4.17) yield the following important result:

hQ are usually small (both are proportional to 1/r), we

shall examine an expansion of det (G ) w.r.t hQ and Result 2. The gain of the maximal bearing-rate change
# 
hQ around the point (0, 0). Then, we obtain the following (mid-course) is

types of fourth-order expansions (in hQ and hQ ) of
  det [Fl/2, l/2 (x,!x)]
det (G ), given by Eqs. (4.16) and (4.17): K134 l\ x\
# det (Fl (x))
(det G )KK(y(x!y)),
# K134 (*x)\, (4.20)
(4.18)
K(x(x!y))
—————
 The type of the expansion only depends on the relative values —————
of i , i , i , i .
     l measurements associated with hQ , i"1, 2.
G G
J.-P. Le Cadre, S. Laurent-Michel/Automatica 35 (1999) 591— 606 601

where *x denotes the total bearing variation (i.e. optimized by means of a standard numerical optimiza-
*x"lx). For usual scenarios, *x is (quite) small with tion procedure. The optimized observer trajectories are
regard to 1 and, therefore, the increase in the FIM deter- presented in Fig. 5. Note that the general shape is a ‘‘Z’’
minant gained by optimized observer maneuvers may be which is stretched when the source-observer distance
rather impressive. Further, note that this gain is propor- decreases. This result is quite general, whatever the rela-
tional to (*x)\. The above calculation is easily extended tive source-observer position and the number of observer
to the case of a maneuvering source. The dimension legs.
of the state vector is then equal to 6, while the gain The computation of the bearing rates is also quite
of a bearing-rate change is, this time, proportional illustrative (see Fig. 6). Actually, we see that the optimal
to (*x)\. headings u induce a bang-bang behavior of hQ . More
I
The following example may be rather enlightening. We precisely, the sequence of optimized controls (i.e. the
present in Fig. 3, the values of det Fl , l (x,!x); observer headings) leads to choose alternatively the two
 
(x"10\) as calculated as a function of the first leg bearing-rate bounds (i.e. !hQ , hQ , see Appendices A,
 
length l (see Appendix C). We can remark that the B and D). This quite agrees with the theoretical results.

optimum is attained for very unequal leg lengths. Actual- The previous results are more formally summarized by
ly, it seems that the optimum corresponds to a ‘‘long’’ the following property.
first leg in order to maximize the observer baseline,
followed by a ‘‘short’’ second leg.
Formula (4.19) may be easily extended to the three-leg
case (i.e. +x, y, z,). As previously, det F l ,l ,l is an homo-
I   
geneous polynomial in (x, y, z), i.e.

 
1
det Fl K P (l , l , l ) xGyHzI , (4.21)
,l ,l
   (pr) G H I   
G H I
with

04+i, j, k,44 and i#j#k"4.

For the sake of brevity, the analytical expressions of the


P are not detailed. Practically, for equal legs (i.e.
G H I
l "l "l "l), the maximum value (x"!y"z) of
  
det F l ,l ,l is approximately 45 lhQ .
I   
Let us now illustrate the above results. The source-
observer scenario is depicted in Fig. 4. The source is in
rectilinear and uniform motion. The observer motion is
modelled by a three-leg path. Furthermore, it is assumed Fig. 4. Optimized receiver trajectory for two source-receiver ranges.
that the modulus of the observer velocity is constant. The
controls are the successive observer headings. They are

Fig. 3. Evolution of the FIM determinant as a function of the first leg


length for a two-leg receiver trajectory. Fig. 5. Details of optimized receiver trajectories.
602 J.-P. Le Cadre, S. Laurent-Michel/Automatica 35 (1999) 591— 606

Fig. 6. Values of hQ for the two scenarios. Fig. 7. Values of hQ for the two scenarios.

Proposition 5. ¸et x , x , 2 , x be the consecutive 4.3. Geometric interpretations of the properties of the FIM
  L
bearing rates, then determinant
arg , 2 , x max det (FIM)"e(hQ ,!hQ , hQ , ),
V L    2 Since we are especially interested in the effects of ob-
e"$1. (4.22) server maneuvers, we shall investigate them by means of
the previous results and differential calculus. Consider,
We refer to Appendix D for a proof of Proposition 5. for instance, the following determinant (E"E ):

Since the values of (!hQ , hQ ) can be estimated by the
  f (y)"det (E, RG E, RH E, RI RJ E), (4.24)
observer (e.g. from the estimated bearings), it remains to  V  V  V  W
determine the optimal number of switching (Sussmann, where x"hQ , y"hQ .
1979) (from hQ to !hQ ) as well as their locations.  
  Let us now calculate the partial derivatives *f/*y(x); we
Using the previous results, the problem may be for-
mulated as follows. Consider a multileg observer obtain
trajectory, then the problem consists in maximizing
*f
det (Fl ,l ,l , 2 ) (l is the length of the ith leg) given below: (x)"l det (E, RG E, RH E, RI>J\ S E), (4.25)
   G *y  V  V  V  V
(pr)det (Fl ,l ,l )+P(l ,hQ )hQ #P(l , l , hQ ) (2hQ )
         where S "(*/*y)R ) ,, or, explicitly
#P(l , l , hQ ) (2hQ )#2 . (4.23)  V  W WV
   
 
S 0
The polynomials P(l , l ) have the form given in Ap- S "  V , (4.26)
G H  V S S
pendix C. The corresponding optimization problem may  V  V
be solved by numerical methods. However, it seems with
rather impossible to derive a general bound relative to

 
the number of switching. !sin x !cos x
S " .
Finally, if the cost functional is replaced by the trace,  V cos x !sin x
the (optimized) observer trajectory is quite different.
Using the definitions of R and S , and denoting
Actually, minimizing the distance becomes quite pre-  V  V
dominant. Thus, the evolutions of the bearing rates J the two-dimensional n/2 rotation matrix, elementary
corresponding to the scenario depicted in Fig. 4 are calculations yield:
presented in Fig. 7. We may note that the optimized
bearing rates no longer exhibit a bang-bang behavior. Proposition 6. ¹he following properties hold:
A comparison of the statistical bounds, associated with *
the observer trajectories corresponding to the optimiza- S "JR , S "JR "!R ,
 V  V *x  V  V  V
tion of the two functionals (i.e. det and tr), shows a very (4.27)
clear supremacy of the det. RI SIY "SIY RI "SI>IY, RI SIY "SI>IY .
 V  V  V  V  V  V  W  V>W
J.-P. Le Cadre, S. Laurent-Michel/Automatica 35 (1999) 591— 606 603

The above results give the algebraic rule for analyzing This approximation is valid as long as rR /r1. If this
the partial derivatives of f. Thus, from Proposition 6, we assumption is valid and if the total duration is ‘‘reason-
directly deduce able’’, the conclusions of the previous sections still hold,
since the effects of bearing-rate changes are preeminent.
*f
(x)"l det (E, RG E, RH E, SI>E),
*y  V  V  V
(4.28) 5. Conclusion
*f
(x, x)"!l det (E, RG E,RH E, RI>J E), etc.
*y  V  V  V The optimization of the observer maneuvers has been
considered along this paper. This problem is not relevant
At this point, it is worth noting that the vector
of classical optimal control. So, a large part of this paper
SK E"(!sin mx, cos mx,!m sin mx, m cos mx) *
 V is centered around the approximation of the cost func-
(Eqs. (4.25) and (4.26)) is approximately orthogonal to the
tional. Using basic tools of multilinear algebra, it has
vectors +E, RG E, RH E,. This fact is typical of a
 V  V been proved that this functional may be accurately ap-
four-dimensional state vector and corresponds to a
proximated by a functional involving only the successive
diversity in maneuvers. Thus Eq. (4.28), (*f/*y)(x)
bearing-range rates of the source. The approach is thus
is proportional to x, while (*/*y)(x, x) is proportional
quite indirect but has the great advantage to involve only
to x, so that
observed data. In particular, it has been shown that
f (y)+cx(y!x). (4.29) under the long-range and bounded controls hypotheses,
the sequence of optimal control lies in the general class of
From Eq. (4.28) it is clear that the increase of det (F ) is bang-bang (relatively to the bearing-rates) one. These
V W
maximized when the terms (x(x!y)) are maximized. results have been illustrated by simulation results. They
Since hQ is bounded, an optimal sequence of controls is demonstrate the interest of maneuver diversity. More
necessarily a bang-bang one (see Appendix D). generally, they provide us with a simple and feasible
approach for optimizing the observer trajectory.
4.4. The effects of range variations

Up to now, the effects of range variations have not Acknowledgements


been considered. However, the analysis is greatly
simplified if we remark that the effects of range and The authors would like to thank the reviewers and the
bearing-rate variations are geometrically decoupled. editor for their many helpful comments.
This follows easily by considering det (G ). Indeed,
#
including the range, the elementary determinant det (G )
#
becomes Appendix A: Calculation of the bearing-rate bounds

 
1 1
det G "det RG E, , RG E , This appendix deals with the calculation of the bounds
# r  2 r 
G G for the bearings rate hQ under the following hypothesis:
1 1 1. The source velocity vector V is fixed.
" 2 det (RG E, 2 , RG E),
 
(4.30) Q
r r 2. The modulus v of the observer velocity is fixed
G G 
(#V #_v ).
so that  
First, recall the basic expression for the bearing-rate hQ :

  
1 1 
det(F l)"(1/p) 2 det(RG E, 2 , RG E) . 1
I r r   hQ " det(V, R),
# G G r
More precise calculations can be achieved if we consider
where V and R are the relative velocity and position
(for instance) a first-order expansion of the source-
vectors.
observer distance (i.e. r "  r #irR ). Roughly, the matrix
I>G I From the above formula hQ is maximum when V is
R is then replaced by (1#rR /r)\R . For instance, the
  orthogonal to R, i.e.
basic result then becomes

  
v a r
1 V" V " W . (A.1)
det(F )J P (hQ )Q (rR ). v r !r
I  pr J J W V

with (4.31)
—————
Q (rR )"r#g r rR #g rrR #g r rR #g rR .  On a leg, the following formula holds: h® "!2 (rR /r) hQ .
J      
604 J.-P. Le Cadre, S. Laurent-Michel/Automatica 35 (1999) 591— 606

We thus have easily deduced:

 
#V#"a"#V ##v!2V*V *H 2v v
 Q  Q jQ "! "! sin(u!h) cos(2h!u)#sin(h) ,
 *r r r
"!v#v#2V*V
 Q  V

 
a *H 2v !v
jQ "! "! sin(u!h) sin(2h!u)#cos(h) .
"(v!v )#2 V* JR,
Q   *r r r
r  W
where J is the n/2 2D-rotation matrix. The scalar a is (B.4)
thus defined as one of the solutions of the second-order Thus, the time derivatives of the costates j and j
equation  
satisfy the following equation:
2 hQ rR
a! (V* JR)a#(v!v )"0. (A.2) jQ r !jQ r "!2 . (B.5)
r   Q W V r
Expliciting the scalar product V* JR, i.e. Then, integrating by parts the optimality condition

(V* JR"ar#V* JR), the previous equation becomes
  *H/*u*"0, we obtain
V* JR j cos h!j sin h"0. (B.6)
!a!2a  #(v!v )"0. (A.3)  
r  Q
It remains to determine the costates j and j . Differ-
 
If the modulus of the observer velocity (v is superior entiating (versus time) Eq. (B.6) yields

to the source one (v ), the above equation has two roots, jQ cos h!jQ sin h!(j sin h#j cos h)hQ "0.
Q
a and a , of opposed signs and the following inequality    
 
holds: Now
!hQ 4hQ 4hQ , hQ rR
  jQ cos h!jQ sin h"!2 ,
  r
where (A.4)
so that, (B.6):
hQ "a , !hQ "a .
    rR rR
j "!2 sin h and j "!2 cos h. (B.7)
 r  r
Appendix B: An optimal control approach From Eq. (B.7), we immediately deduce that

This appendix deals with the optimal control problem vh rR  2


jQ "!2 cos (2h!u)#8 sin h# uR hQ sin h. (B.8)
of the Section 3.2 and, more precisely, with the solution of  r r r
the optimization problem associated with (3.7). Collecting Eqs. (B.8) with (B.4), yields
Expliciting the PMP optimality conditions, we have:


rR 


*H * hQ  4 #uR hQ "!1. (B.9)
"! #j v cos u!j v sin u, r
*u *u r  

 
*H 1 *hQ  *r
"! r! hQ  . (B.1) Appendix C: The bearing-rate change polynomials
*r r *r *r
V V V
Owing to the classical relations (v v )hQ " Explicit expressions of the polynomials +P (l , l ),
  G   G
v/r sin(u!h), rR "v cos(u!h), *H/*u is easily calculated are given below: 
yielding:
P "(a\)l (n!l )[5l#46ln!57l n#32ln
*H rR hQ       
"!2 #j v cos u!j v sin u. (B.2) !113l n#102l n!19l n#4l n],
*u r      
P "(b\)l (n!l )[!10l!56ln#87l n
Using classical differential calculus, *hQ /*r and *hQ /*r      
V W
are explicited: !42ln#108l n!102l n#19l n!4l n],
    
*hQ v P "(a\)l [30l!270ln#480ln!405l n
     
"! cos(2h!u),
*r r #360l n!316ln#144ln!27ln#4l n],
V (B.3)     
*hQ v
" sin(2h!u).
*r r —————
W  These results are obtained by means of symbolic computations.
From Eqs (B.1) and (B.3), the following equalities are  n is the total number of measurements.
J.-P. Le Cadre, S. Laurent-Michel/Automatica 35 (1999) 591— 606 605

P "(b\)l [!10l#36l n!9l n!48l n Now, both a and b are negative, so is the term
     
(a/b)(b#a). Thus, the minimum of f (x, y) cannot be
#36l n#5l ],
  inside the square of constraints which implies that the
P "(a\)l [5l!36l n#54l n!32l n#9l n], minimum of f (x, y) can be achieved only on the bound-
      
(C.1) ary of the constraint domain and more precisely for an
where extreme point. Thus, the function f (x, y) is minimum for
x"1, y"!1. Proposition 5 is thus proved in this case.
a"25920, b"12960, n"l #l . Extension to the general case is a bit more intricated,
 
Using Eq. (C.1), the following equality is easily shown: even if the basic idea is very similar. Consider, for in-
stance, the case of three bearing-rates (x, y, z) and the
 following functional f (x, y, z):
P (l , l )x\GyG\"0 ∀x, y"x, l , l . (C.2)
G    
G f (x, y, z)"ax(y!x)#ay(y!x)#by(z!y),
#bz(z!y)#cz(z!x)#cx(z!x).
Appendix D: A proof of the bang-bang property
We are now dealing with the following optimization
The aim of this section is to prove Proposition 5. For problem:
the clarity of presentation, the analysis will be first re- min f (x, y, z)
stricted to the case of a unique bearing-rate change.
Then, it will be extended to the general case. x!140, y!1, z!140,
(D.4)
In the unique bearing-rate change case (x and y are the !x!140, !y!1, !z!140
consecutive bearing rates), the maximization of the FIM a, a, 2 , c, c(0.
determinant reverts to the following optimization prob-
lem (see Eqs. (4.15)—(4.17)): For an interior point, the Lagrangian stationarity then
implies
min f (x, y)
*f *f *f
f (x, y)"ax(y!x)#by(y!x), # # "0,
*x *y *z
x!140, y!140, (D.1)
or, explicitly
!x!140, !y!140
[a(y!x)#c(z!x)]x#[a(y!x)#b(z!y)]y
a and b(0.
#[c(z!x)#b(z!y)]z"0,
The Lagrangian of the above problem is then
whence
¸(x, y)"f (x, y)#u (x!1)#u (!x!1)
 
z"d x#d y,
#v (y!1)#v (!y!1).  
 
with (D.5)
Assume that the solution to the problem lies (strictly)

 
inside the square of constraints, then the stationary a(y!x)#c(z!x)
d "! ,
(Kuhn—Tucker) conditions imply  c(z!x)#b(z!y)
*f
 
a(y!x)#b(z!y)
*x
"2ax(y!x)!2(ax#by)(y!x)"0, d "! .
 c(z!x)#b(z!y)
(D.2)
*f From Eq. (D.5), we immediately deduce that d and d
*y
"2by(y!x)#2(ax#by)(y!x)"0.  
are negative. The following expressions of the elementary
In particular, the optimal values of x and y must satisfy terms of f (x, y, z) are then obtained:

*f *f by(z!y)"by[d x#(d !1)y]"f (x, y),


# "2(y!x)[ax#by]"0,   
*x *y cx(z!x)"cx[(d !1)x#d y]"f (x, y),
  
so, that (yOx): bz(z!y)"b(d x#d y)[d x#(d !1)y]"f (x, y),
    
ax#by"0, cz(z!y)"c(d x#d y)[(d !1)x#d y]"f (x, y).
    
whence (D.6)

 
a#b  a The following simple remarks are then instrumental:
f (x, y)" (b#a)x. (D.3)
b b b, b, c, c are altogether negative, as well as d , d , d !1,
  
606 J.-P. Le Cadre, S. Laurent-Michel/Automatica 35 (1999) 591— 606

d !1. Using the previous reasoning, we thus see that the Mc Cabe, B. J. (1985). Accuracy and tactical implications of bearings-
 only ranging algorithms. Oper. Res., 33(1), 94—108.
minimum of + f (x, y), is attained for x"1, y"!1.
G G Nardone, S. C., & Aidala, V. J. (1981). Observability criteria for
Since the same (minimum) property is also verified by the bearings-only target motion analysis. IEEE ¹rans. Aerospace
functions ax(y!x) and ax(y!x), the minimum of Electron. Systems, AES-17 (2) 162—166.
f (x, y, z) is necessary attained on the boundary of the Nardone, S., Lindgren, A., & Gong, K. (1984). Fundamental properties
cube of constraints. Hence we easily deduce that the and performance of conventional bearings-only target motion anal-
minimum of f is attained for an extreme point of the ysis. IEEE ¹rans. Automat. Control, AC-29(9), 775—787.
Olsder, G. J. (1984). On the optimal manoeuvering during bearings-
boundary, i.e. x"1, y"!1, z"1. Note that the values only tracking. Proc. 23rd Conf. on Decision and Control
of f (x, y, z) are identical on the points x"1, y"!1, (pp. 935—940). Las Vegas, NV, December.
z"1 and x"!1, y"1, z"!1. Passerieux, J. M., & Van Cappel, D. (1998). Optimal observer maneuver
By recursion, this reasoning is easily extended to for bearings-only tracking. IEEE ¹rans. Aerospace Electron.
higher dimension. Systems, 34(3), 777—789.
Passerieux, J. M., & Van Cappel, D. (1991). Manoeuvre optimale en
trajectographie passive a partir d’azimuts, Actes du 13-ème colloque
Gretsi (pp. 285—288). Juan-les-Pins, September.
Sussmann, H. J. (1979). A bang-bang theorem with bounds on the
References number of switchings. SIAM J. Control Optim. 17(5), 629—651.
Teo, K. L., Goh, C. J., & Wong, K. H. (1991). A unified computational
Aidala, V. J., & Hammel, S. E. (1983). Utilization of modified polar approach to optimal control problem, Pitman Series in Pure and
coordinates for bearings-only tracking. IEEE ¹rans. Automat Applied Math., Longman Scientific and Technical.
Control, AC-28, 283—294. Trémois, O., & Cadre, J. P. (1996). Target motion analysis with multiple
Atwood, C. L. (1973). Sequences converging to D-optimal designs of arrays: Performance analysis. IEEE ¹rans. Aerospace Electron.
experiments. Ann Statist., 1(2), 342—352. Systems, AES-32(3), 1030—1045.
Barniv, Y. (1985). Dynamic programming solution for detecting Whittle, P. (1973). Some general points in the theory of optimal
dim moving targets. IEEE ¹rans. Aerospace Electron. Systems, experimental design, J. Roy. Statist. Soc. Ser. B, 35, 123—130.
AES-21(1). Yokonuma, T. (1992). ¹ensor spaces and exterior algebra. Transl. of
Blom, H. A. P., & Bar Shalom, Y. (1988). The interacting multiple Math. Monographs (Vol. 108), American Mathematical Society,
model algorithm with Markovian switching coefficients. IEEE Providence, RI.
¹rans. Automat. Control, AC-33(8), 780—783.
Chang, C. B., & Tabaczinski, J. A. (1984). Application of state
estimation to target tracking. IEEE ¹rans. Automat. Control,
AC-29(2), 98—109. Stephane Laurent-Michel received his en-
Darling, R. W. R. Differential forms and connection. Cambridge, UK: gineering diploma in signal and image
Cambridge University Press. 1994. processing from ENSEEIHT and a post-
Fawcett, J. A. (1988). Effects of course maneuvers on bearings-only graduate diploma from the Institut Na-
range estimation. IEEE ¹rans. Acoust. Speech Signal Process. 36(8), tional Polytechnique de Toulouse (INPT)
1193—1199. in 1996. He also received a Master degree
Hammel, S. E. (1988). Optimal observer motion for bearings-only in Electronics for telecommunication from
the Ecole Nationale Superieure des Tele-
localization and tracking, Ph.D. thesis, University of Rhode Island.
communications (ENST Paris) in 1997. In
Hammel, S. E., Liu. P. T., Hilliard, E. J., & Gong, K. F. (1989). Optimal 1997 he developed a digital narrow band
observer motion for localization with bearings measurements. received from Schlumberger. He is now an
Comput. Math. Appl. 18 (1—3), 171—186. application engineer with SYNOPSYS
Helferty, J. P., & Mudgett, D. R. (1993). Optimal observer trajectories Inc., Mountain View, CA working on COSSAP in the systems group.
for bearings-only tracking by minimizing the trace of the His areas of interest are digital signal processing and telecommunica-
Cramel—Rao bound. Proc. 32nd Conf. on Decision and Control tion systems.
(pp. 936—939). San-Antonio. TX, December.
Hocking, L. M. (1997). Optimal control. An introduction to the theory
with applications. Oxford Applied Math. Series. J. P. Le Cadre received the M.S. degree in
Holtsberg, A. (1994). A statistical analysis of bearings-only tracking, pure Mathematics in 1977, the ‘‘Doctorat
Ph.D. thesis, Department of Math. Statistics, Lund Institute of de 3\CKC cycle’’ in 1982 and the ‘‘Doctorat
Technology. d’Etat’’ in 1987, both from INPG,
Grenoble. From 1980 to 1989, he worked
Jauffret, C., & Musso, C. (1991). New results in target motion analysis. at the GERDSM (Groupe d’Etudes et de
UDT Conf. (pp. 239—244). Paris, 23—25 April. Recherche en Detection Sous-Marine),
Kershaw, D. J., Evans, R. J. (1994). Optimal waveform selection for a laboratory of the DCN (Direction des
tracking systems. IEEE ¹rans. Inform. ¹heory, I¹-40(4), 1536—1551. Constructions Navales), mainly on array
Lancaster, P., & Tismenetsky, M. (1985). ¹he theory of matrices and signal processing. In this area, he con-
(2nd ed.) New York: Academic Press. ducted both theoretical and practical re-
Le Cadre, J. P., & Jauffret, C. (1997). Discrete-time observability and searches (towed arrays, high resolution
estimability analysis for bearings-only target motion analysis. IEEE methods, performance analysis, etc.) and has been responsible of
¹rans. Aerospace Electron. Systems, AES-33(1). 178—201. a working team (GdR, CNRS). Since 1989, he is with IRISA/CNRS,
where is Directeur de Recherche at CNRS (National Center for Scient-
Le Cadre, J. P., & Trémois, O. (1997). The matrix dynamic program- ific Research). His dimains of interest have now moved towards other
ming property and its implications. SIAM J. Matrix Anal. 18(4), topics like system analysis, detection, multitarget tracking, data associ-
818—826. ation and operations research. He has received (with O. Zugmeyer) the
Liu, P. T. (1988). An optimum approach in target tracking with bear- Eurasip Signal Processing best paper award (1993) and is a member of
ings measurements. J. Optim. ¹heory Appl. 56(2), 205—214. IEEE.

You might also like