Adaptive Mobile Sensor Positioning for Multi-Static Target Tracking-4rV.pdf

(1385 KB) Pobierz
641081934 UNPDF
I. INTRODUCTION
Adaptive Mobile Sensor
Positioning for Multi-Static
Target Tracking
Recent improvements in battery, microcontroller,
and sensor technologies have resulted in the
development of autonomous vehicles that are
inexpensive, dependable, and simple to operate.
Unmanned air vehicles (UAVs) have well-known
defense applications and are useful in many civilian
applications as well, such as border and coast patrol,
fire perimeter monitoring, search and rescue, etc.
UAVs are particularly well suited for situations that
are too dangerous for direct human monitoring. While
a single autonomous UAV can provide information
that is otherwise unattainable, cooperation among
a team of UAVs can dramatically improve sensing
performance [1—5]. The mobility of each UAV within
the team can be exploited to adaptively reconfigure
the sensing array in response to the environment and
the actions of the object being sensed [6—8]. It is this
additional degree of freedom we wish to exploit in a
multistatic radar tracking scenario.
Using mobile and controllable sensing platforms
is a relatively new area of research, but has attracted
considerable attention in the generic sensor network
literature. This type of problem is sometimes referred
to as sensor management. The majority of work in
sensor management for target tracking assumes one of
four practical observation models: range only [9, 10],
bearing only [11, 12], range/bearing [2, 13], and
more recently, range/range-rate/bearing [14]. These
observation models are common due to the prevalent
use of directional laser and radar sensors. However,
small UAVs may have a difficult time making accurate
bearing measurements given their limited aperture
and the fact that they are much more sensitive to
wind-blown disturbances. Moreover, power constraints
on small UAVs limit the range and strength of their
radar transmissions.
Other papers have focused on static sensors
that are capable of toggling between active and
resting states [15—19] or arranging the scheduling
and reporting of measurments in an optimal manner
[20, 21]. The placement of static sensor nodes for
tracking in an underwater scenario is investigated in
[3], [22], [23]. The advantages of mobile sensors are
studied in [6], [24], where sensors are constrained to
remain in small groups that retain observability of the
target states. The motion and geometric formation
effects of these groups on tracking performance are
then investigated. In [25], the optimal positions of
a set of UAVs in a tracking scenario are determined
based on the Cramer-Rao bound (CRB). These results
provide insight into the relationship between the
angular distribution of the UAV team and target
state estimation using range, range-rate, and bearing
measurements. However, the results are derived
under the assumption that the accuracy of the UAV
measurements are constant, and not a function
PENGCHENG ZHAN
DAVID W. CASBEER
Brigham Young University
A. LEE SWINDLEHURST, Fellow, IEEE
University of California at Irvine
Unmanned air vehicles (UAVs) are playing an increasingly
prominent role in both military and civilian applications. We
focus here on the use of multiple UAV agents in a target tracking
application where performance is improved by exploiting
each agent’s maneuverability. Local time-delay and Doppler
measurements made at each UAV are used as inputs to an
extended Kalman filter (EKF) which tracks the target’s position
and velocity. Two simple metrics are defined to quantify the
accuracy of the tracking algorithm, and heading feedback to
the UAVs is used to minimize the metric and improve tracking
performance. A simplified version of one of the algorithms that
reduces computational complexity is also presented. Simulations
demonstrate the significant improvement that results when
the UAV sensors are allowed to be optimally positioned during
tracking.
Manuscript received June 20, 2007; revised March 31, 2008;
released for publication September 2, 2008.
IEEE Log No. T-AES/46/1/935931.
Refereeing of this contribution was handled by W. Koch.
This work was supported by the U.S. National Science Foundation
under Information Technology Research Grant CCF-0428004.
Authors’ current addresses: P. Zhan, ArrayComm LLC, San Jose,
CA; D. Casbeer, Dept. of Electrical and Computer Engineering,
Brigham Young University, 459 Clyde Building, Provo, UT 84602,
E-mail: (casbeer@byu.edu); A. L. Swindlehurst, Dept. of Electrical
Engineering and Computer Science, University of California at
Irvine, Irvine, CA.
0018-9251/10/$26.00 ° 2010 IEEE
120
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 11:53:34 UTC FROM IEEE XPLORE. RESTRICTIONS APPLY.
641081934.004.png
of the relative position of the UAVs and the target.
In addition the solution proposed in [25] provides an
optimal set of UAV positions, but no path planning for
the UAVs or any guarantees that it is possible for the
dynamically constrained UAVs to reach the resulting
positions in the allotted time.
In this work, we assume a multistatic radar
scenario in which a transmitter, either airborne or on
the ground, illuminates a target with a radar signal.
A team of UAVs receives the target reflection, and
each UAV makes a local time-delay and Doppler
measurement. These measurements are then sent
to a centralized processor or base which uses this
collected data set to track the target’s position and
velocity. Based on the calculated accuracy of the
tracking algorithm, the base determines trajectories
for the UAV team in order to optimize performance
and sends heading commands to the UAVs to
best position them for the next measurement. The
maneuverability of the UAVs provides the base with
the ability to improve tracking performance through
the tunable parameters in the tracking algorithm.
We present two general optimization algorithms
for the problem, and a closed-form solution that
is considerably simpler to implement but performs
essentially as well as the optimal approaches in many
situations.
The format of the paper is as follows: In
Section II we describe the problem setup, the
assumed dynamic models, and relevant aspects
of the extended Kalman filter (EKF). The issue
of range-dependent measurement accuracy is also
addressed. In Section IIIA and IIIB, two algorithms
for determining the optimal UAV heading angles
based on the EKF updates are presented. A more
computationally efficient approach is described in
Section IIIC, based on heuristic observations of one of
the algorithms in simple scenarios. Simulation results
are presented in Section IV, and some conclusions are
drawninSectionV.
Fig. 1. Multiple UAVs tracking single target.
The UAVs are assumed to be capable of velocity
and altitude hold with an autopilot such as the one
described in [26]. Therefore, for simplicity, we
consider only a two-dimensional scenario. We define
the position of the base station as the origin for our
( x , y ) coordinate system where x and y correspond to
the cardinal directions East and North, respectively.
Let x , y , V x ,and V y represent, respectively, the position
of the target and the components of the target velocity
vector in the ( x , y ) directions. Similarly, let x i and y i
denote the position of the i th UAV, and assume that
each UAV is flying at the same constant speed V ,with
heading angle à i measured counterclockwise from
the x axis. According to [27] and after appropriate
scaling, the time-delay and Doppler measurements can
be written as
¿ i = 1
Μ p
q
x 2 + y 2 +
( x¡x i ) 2 +( y¡y i ) 2
(1)
c
Ã
!
f i = 1
V x x + V y y
p
V p
x 2 + y 2 +
p
(2)
¸
( x¡x i ) 2 +( y¡y i ) 2
where
¡V sin à i )( y¡y i ) :
In the above equation, c and ¸ stand for the speed of
light and the wavelength of the signal, respectively.
The problem we address in this paper is how
the mobility of the UAV receivers can be exploited
to improve the system’s tracking performance. In
our solution to this problem, the time-delay and
Doppler estimates collected by the UAV team are
uploaded to the base and used as measurement
updates in an EKF tracking the target. Simultaneously
optimal heading commands for each UAV are
calculated at the base station and then transmitted to
the agents. The precise optimality criteria used for
computing the heading commands is presented in
Section III.
¡V cos à i )( x¡x i )+( V y
II. BACKGROUND AND SYSTEM DEFINITION
A. Problem Statement
As illustrated in Fig. 1, we consider a multistatic
radar problem in which a set of N UAVs receives the
signal reflected by an airborne target illuminated by a
base transmitter. We assume that the UAVs are able to
form time delay ( ¿ i ) and Doppler ( f i )estimatesbased
on these signals, 1 and transmit these measurements
back to the base.
B. UAV Receivers and Measurement Error
1 Another viable model would be to use the time difference
of arrival rather than time delay. Using this model the control
algorithms developed later would remain unchanged since the
observation Jacobians with respect to target parameters are identical
under both models.
The accuracy of the UAV measurements is a
function of the transmitted radar signal, the signal
strength, the length of the propagation path from
transmitter to receiver, and so on. To quantify the
ZHAN ET AL.: ADAPTIVE MOBILE SENSOR POSITIONING FOR MULTI-STATIC TARGET TRACKING
121
AUTHORIZED LICENSED USE LIMITED TO: IEEE XPLORE. DOWNLOADED ON MAY 13,2010 AT 11:53:34 UTC FROM IEEE XPLORE. RESTRICTIONS APPLY.
V p =( V x
641081934.005.png 641081934.006.png
We represent the target state vector at time k as
x k =[ xyV x V y ] T , and define the vector h k ( x k )by
arranging the time-delay and Doppler equations from
each UAV as
h k ( x k )=[ ¿ 1
¢¢¢¿ N
f 1
¢¢¢f N ] T
Fig. 2. Bistatic radar scenario.
where, as described earlier, N is the number of UAV
receivers. Using a linear discrete-time system and an
additive noise observation model, the system is then
written as
time-delay and Doppler measurement accuracy for
UAV i , we use the CRB as derived by [28]:
x k = A k x 1 + ! 1
(6)
Μ
¶·
¸
= 2 E r
N 0
E r
E r + N 0
! 2 !t
!t t 2
CRB
¡ 1
i
(3)
z k = h k ( x k )+ º k
(7)
where
Z
where A k is the state transition matrix, ! 1
»
! 2 = 1
2 ¼
1
¡1 ! 2 jS ( j! ) j 2 d!
»N ( 0 , R k ) are, respectively, the
process and measurement noise. Both are assumed
here to be uncorrelated Gaussian random vectors.
As with the standard Kalman filter, the EKF
can be divided into two stages: the time update and
measurement update steps. With the time update, we
propagate the target’s state estimate and prediction
covariance matrix as
Z
¡1 u ˜ s ( u ) @ ˜ s ¤ ( u )
1
!t =Im
@u du
Z
1
¡1 u 2 j ˜ s ( u ) j 2 du
! i =2 ¼f i
˜ s ( u )and S ( j! ) are, respectively, the complex envelope
of the transmitted signal and its Fourier transform,
Im represents the imaginary part of its argument, E r
is the average received signal power, and N 0 = 2isthe
spectral density of the white bandpass Gaussian noise
from the receiver on each UAV. Inverting (3) yields
the following explicit expression for the CRB:
t 2 =
x
k = A k x 1
(8)
P
k = A k P 1 A k + Q 1 :
(9)
Once we have the measurements from the sensors, we
update our estimate as
x k = x
k + K k ( z k
¡ h k ( x
k ))
(10)
·
¸
N 0 ( E r + N 0 )
2 E r [ ! 2 t 2 ¡ ( !t ) 2 ]
t 2 ¡!t
¡!t ! 2
CRB i =
: (4)
P k =( I ¡ K k H k ) P
¡
k
(11)
where
The accuracy of the second step, finding the
position and velocity estimates for the target from the
time-delay and Doppler estimates, are quantified in
the next section by the EKF. This is done through the
Jacobian H k (defined subsequently) which maps the
accuracy of the time-delay and Doppler estimates to
the position and velocity of the target.
We use the standard bistatic radar equation [29]
to model the received signal power E r [30] based on
the geometry of the bistatic radar scenario depicted in
Fig. 2:
K k = P
k H k ( H k P
k H k + R k ) ¡ 1 : (12)
In the above equations, H k =( @ h k =@ x ) T is a function
of the UAVs’ positions and headings.
The basic idea of this paper is to reduce the size
of the estimation error by updating the heading of the
UAVs, and hence H k . In essence, we are applying an
outer control loop around the Kalman filter.
D. System Dynamics
E r = E t G t G r ¸ 2 ¾ b
(4 ¼ ) 3 R b R i
(5)
The approach presented in the next section is
general enough to accommodate any particular model
for A k , Q k ,or R k . In the simulation studies presented
later, we assume a simple constant velocity target
model described by the state transition matrix
where E t is the transmit power; G t , G r are antenna
gains at the base station and UAV side, respectively;
¸ is the wavelength of the transmitted radio signal;
and ¾ b is the radar cross section (RCS) of the target.
2
4 10 ¢ t
0
3
C. Extended Kalman Filter
A k =
01 0 ¢ t
00 1 0
5
(13)
The need for the EKF is apparent from the
nonlinear measurement equations in (1) and (2).
00 0 1
122
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 11:53:34 UTC FROM IEEE XPLORE. RESTRICTIONS APPLY.
N ( 0 , Q 1 )and º k
¡
¡
¡
¡
¡
¡
641081934.007.png 641081934.001.png
where ¢ t is the time interval between samples. We
also use a simple white process noise model:
nonlinearities are accounted for through the Jacobian
H k in the EKF.
Q k =diag( ¾ x , ¾ y , ¾ V x , ¾ V y ) :
(14)
III. HEADING ALGORITHMS
The definition of the measurement noise model
is more critical to the problem at hand, since it will
be a function of the UAV positions. Assuming that,
given the current target state, the time-delay and
Doppler estimates at each UAV are independent from
the estimates made by the rest of the UAV team,
the expression for the CRB derived above can be
used to find the elements of the measurement noise
covariance. For 1 ·i·N ,
In this section we present two algorithms for
finding the optimal UAV heading commands from
the EKF. In the first algorithm, the headings are
chosen to minimize the weighted trace of the EKF
prediction error covariance matrix one step into
the future. In the second, UAV headings are found
that maximize the amount of new information
provided by the measurements at the next step. A
heuristic simplification of the second algorithm is also
presented that leads to a closed-form solution to the
problem and considerable computational savings.
<
CRB i (1,1) j = i
1
2 ¼ CRB i (1,2) j = i + N
R k ( i , j )= E ( º i º
j )=
:
A. Error Covariance Minimization
0
e l s e w h e r e :
When N<i· 2 N ,
The motivation underlying the approach presented
here is simply to fly the UAVs in directions at
time t = k such that the estimation error at time
t = k +1 is made as small as possible. To this end,
we minimize the weighted trace of the one-step
prediction covariance P
<
:
Μ
1
2 ¼
2
CRB i¡N (2,2) j = i
R k ( i , j )= E ( º i º ¤
1
2 ¼ CRB i¡N (2,1) j = i¡N
¡
k +1 [34]. In practical scenarios,
the UAV dynamics limit the rate at which its heading
may change. Therefore, for each step we restrict the
UAVs’ position to an arc defined by the previous
heading. Mathematically, the idea is to find the vector
of heading commands ª k =[ Ã 1, k :::Ã N , k ] T as follows:
0
e l s e w h e r e :
We assume the ability of the base to pass a
heading command à i , k to UAV i .Assumingan
instantaneous response to heading commands, the
UAVs will fly in a straight line with a velocity V
and heading à i during each time interval. The inertial
position of the UAV i at time k can be approximated
by
ª k =argmin
ª k
tr( W k P
k +1 )
(17)
i , k
¡Ã i , 1
j·C
8
i =1 :::N
x i , k = x i , 1 + V cos( Ã i , k ) ¢ t
(15)
where W k is a positive semidefinite weighting matrix
and C is the upper bound on the turning rate of the
UAVs.
Replacing P
y i , k = y i , 1 + V sin( Ã i , k ) ¢ t : (16)
This approximation will be valid as long as the new
heading commands are not too different from the
UAV’s current heading, and the time steps are short
enough.
It is known that due to the geometry of a bistatic
radar system, the bistatic radar ambiguity function,
which is an indicator for how well position and
velocity can be determined from the received
waveform, is a nonlinear function of the transmitter,
receiver, and target positions [31—33]. In our system,
each UAV calculates only the time-delay and Doppler
information from the received waveform that is
possible to calculate under the assumption that the
received signal can be unambiguously associated with
the target being tracked. The time-delay and Doppler
measurements are then passed to the base station for
processing; the nonlinearities arising in the bistatic
radar ambiguity function are encountered at this step
when the velocity and position estimates are extracted
from the measured time delay and Doppler. These
¡
k +1 in (17) with the expressions in (9)
and (11) yields
k H k ] ¡ 1 A k +1 + W k Q k ) :
(18)
Since the dependence of the cost function on the UAV
headings is not explicit, it is useful to examine the
optimization problem in slightly more detail. Equation
(17) depends on the heading angles through H k and
R k in (18), which in turn depend on the positions
of the UAVs at time k , which are a function of their
positions at time 1 and the heading angles chosen
at time 1 (as described by the model in (15) and
(16)). Thus, when we minimize (17) with respect to
ª k , we are minimizing it with respect to the headings
chosen at time 1 after the measurement at time
1, which headings are also the ones in force at
time k before measurement k is taken.
When the number of UAVs is not too large, a
finite grid search can be used to optimize the cost
k ) ¡ 1 + H k R
¡ 1
ZHAN ET AL.: ADAPTIVE MOBILE SENSOR POSITIONING FOR MULTI-STATIC TARGET TRACKING
123
AUTHORIZED LICENSED USE LIMITED TO: IEEE XPLORE. DOWNLOADED ON MAY 13,2010 AT 11:53:34 UTC FROM IEEE XPLORE. RESTRICTIONS APPLY.
¤
j )=
¡
¡
tr( W k A k +1 [( P
641081934.002.png
function. For higher dimensional problems, we
employ a Quasi-Newton approach to minimize (18).
At each iteration the heading is updated by ª m +1 =
ª m
the covariance measurement update in (11), yielding
P
k =( P
k )
¡ 1 + H k R
k H k
(20)
¡ m and g m are the Hessian matrix
and gradient vector of the cost function evaluated at
ª m . Due to the complexity involved in calculating the
Hessian matrix, the Quasi-Newton approach uses an
estimate of the Hessian based on the gradient vector.
The gradient with respect to the heading of the i th
UAV at time k is given by
@ tr( W k P
¡ S
¡ m g m ,where S
where P ¡ 1
k
is the information matrix at step k .The
k H k represents the increase in information
provided by the UAVs, assuming that their headings
are chosen to yield H k . If the UAVs provided no
new information at step k ,then H k R ¡ k H k =0,and
the information matrix would be equal to its value
after 1steps: P ¡ k =( P ¡
¡ 1
Μ
k ) ¡ 1 . The idea, then, is to
k +1 )
W k A k +1 B ¡ 1 @ B
k H k ismaximizedinsome
sense. This approach makes sense intuitively: one can
think of this as choosing a set of vectors H k that lie as
much as possible in the eigenvector space of R ¡ 1
¡ 1
= ¡ tr
i , k B ¡ 1 A k +1
i , k
where
(19)
k with
the largest eigenvalues (or, equivalently, in the space
of eigenvectors of R k with the smallest eigenvalues).
In other words, we try to position the UAVs so that
the new information that they provide is corrupted the
least by the measurement “noise.” Since H k enters into
this term quadratically, an analysis of the impact of
the UAV headings is much simpler. As discussed in
the next section, this allows for the development of an
efficient algorithm for computing the optimal heading
commands.
Under the maximizing information increase (MII)
approach, the heading commands are found by solving
B =( P
¡
k )
¡ 1 + H k R
¡ 1
k H k
@ B
i , k
= @ H k
Μ
@ H k
T
i , k R
k H k +
i , k R
k H k
¡ H k R
¡ 1
k
i , k R
k H k :
The problem boils down to calculating @ H k =@Ã i , k and
@ R k =@Ã i , k ,where H k is 2 4and R k is 2 2 N .
Minimizing the cost function (17) over the correct
control variable will work generally for any adaptive
sensor network. The weighting matrix W k is helpful
in that is can be used to place importance on specific
variables of interest. In our problem the magnitude for
position values was much greater than the velocity,
and by adjusting the elements of W k ,wewereableto
place suitable errors on each of these terms.
ª k =argmax
ª k
det( H k R
¡ 1
k H k )
(21)
i , k
¡Ã i , 1
j·C
8
i =1 :::N:
Assuming that the measurement errors across the
UAVs are uncorrelated, and after regrouping the
time-delay and Doppler equation vector as h k ( x k )=
[ ¿ 1 f 1
B. Maximizing Information Increase
¢¢¢¿ N f N ] T , the covariance matrix R k will have
a block diagonal structure. Each block along the
diagonal is denoted by R i k , which is equivalent to the
CRB matrix discussed in Section IIB. More algebraic
manipulation leads to the following equality:
We investigate here an alternative approach
derived from examining an expression for the Fisher
information matrix (or inverse covariance matrix).
Our motivation is to find an algorithm for determining
the UAV headings that is computationally simpler to
implement than the error covariance minimization
(ECM) method of the previous section. As we will
see, the UAV heading parameters enter into the
information matrix update in a much simpler way
that lends itself more easily to approximation. A
closed-form solution based on such an approximation
is presented in the next section. A computationally
efficient optimization method was found in [9]
for a similar application that used the trace of the
covariance matrix, but it was for a much simpler
measurement model than assumed here.
Rather than optimizing over the information matrix
itself, in our approach we choose the UAV headings
so that the increase in information provided by the
UAVs at the new locations is maximized [2, 35]. More
precisely, note that the information matrix update can
be found by applying the matrix inversion lemma to
X
' k = H k R
k H k =
H i k ( R i k ) ¡ 1 H i k
(22)
i =1
H k =[ H 1 k ::: H N k ] :
The submatrices, H i k , which constitute the
measurement gradient matrix, H k , are defined in (34)
of the Appendix.
By maximizing the new information (21), the
sensors will position themselves such that the next
measurement contains the most information possible.
Notice that this cost metric does not necessarily try to
gain new information that is less known to the system.
It simply tries to find the measurements with the
most information. This metric is perhaps not optimal,
but as shown in the results section, it yields good
performance results. Again because of the simple form
the cost function (21), it allows a closed-form solution
which is presented in the following section.
124
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 11:53:34 UTC FROM IEEE XPLORE. RESTRICTIONS APPLY.
¡ 1
¡
¡ 1
term H k R
choose H k so that H k R
¡
¡ 1
¡ 1
@ R k
¡ 1
N
¡ 1
641081934.003.png
Zgłoś jeśli naruszono regulamin