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

JOURNAL OF GUIDANCE, CONTROL, AND DYNAMICS

Vol. 37, No. 6, NovemberDecember 2014

Arbitrary-Order Sliding-Mode-Based Homing-Missile Guidance


for Intercepting Highly Maneuverable Targets

B. Kada
King Abdulaziz University, Jeddah 21589, Saudi Arabia
DOI: 10.2514/1.G000001
This paper investigates the development and implementation of a new homing-missile guidance system for
intercepting highly maneuverable targets using arbitrary features of the homogeneous high-order sliding-mode
controllers and observers. The concept of this guidance system involves artificially increasing the orders of controllers
and observers to enhance accuracy and robustness of future interceptors. Two advanced guidance laws, that
is, augmented proportional navigation law and direct collision or hit-to-kill law, are considered to construct
acceleration-based and velocity-based attractive sliding manifolds, respectively; and high-order robust exact differ-
entiators are used to compute high-order time derivatives of these manifolds. Real-time estimation of target maneuver
Downloaded by University of Toronto on November 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.G000001

from measurements using these differentiators is also considered. The resulting guidance laws are implemented in
computer simulation using nonlinear interceptor dynamics and engagement kinematics. The results prove that the
guidance system achieves high-level performance and successfully meets direct collision guidance requirements. In
addition, this system shows good adaptability and strong robustness to target evasive maneuvers, parameter
uncertainties, and measurement noises.

Nomenclature control, artificial intelligence, and variable structure control [59].


A = acceleration, g Many of these solutions were derived by 1) assuming perfect
h = z-axis coordinate, m knowledge of target state and maneuver or estimating them using
M = Mach number conventional observers, 2) considering no measurement noise or
q = pitch rate, rads nonprocessing target information, and 3) using ideal or linearized
R = missiletarget closing range, m interceptor dynamics. Even though the resulting guidance laws are
r = relative degree accurate for nonmaneuvering and moderately maneuvering targets,
tf = time of flight or total time of guidance, s these laws could fail to guarantee stringent performance requirements
V = velocity, ms and ensure stable and agile flight when applied for realistic interception
x = x-axis coordinate, m missions involving highly and randomly maneuverable targets.
= angle of attack, deg On the other hand, as the tracking and interception missions have
= flight-path angle, deg to be performed in realistic scenarios, it is of prime importance to
= control surface deflection, deg accurately estimate the target state and maneuver to ensure target kill.
= line-of-sight angle, deg The key to successful target tracking lies in the ability of observers to
= sliding manifold extract useful information about the target state and maneuver.
= time constant, s Assuming a suitable dynamics model, several estimation algorithms
have been designed and inserted in the guidance and homing loops as
target state and maneuver estimating building blocks, such as Gauss
Markov processes, extended Kalman filter (EKF), multiple-model
I. Introduction methods, and artificial intelligence observers [1020]. To describe

F ROM a defense point of view, the deficiency of existing homing-


missile guidance strategies to guarantee satisfactory endgame
performance against maneuvering targets is essentially due to
real-world scenarios, most of these observers continuously update a
linearization around the current state estimate, use a large number of
filters, or represent the target maneuver as a white noise process,
1) target evasive maneuvers, 2) large errors in target state estimation which could be inappropriate in the case of abrupt target maneuvers
and maneuver tracking, 3) noise contaminating seeker measure- or in the presence of measurement noises [21,22].
ments, and 4) interceptor maneuverability saturation. Recently, and due to their impressive performance and strong
Over the last few decades, great strides have been made in robustness, high-order sliding modes (HOSMs), such as smooth
responding to these challenges and interesting investigations have second-order sliding modes, outer-loop sliding modes, and homoge-
been conducted, leading to new perspectives and solutions. Tradi- neous sliding modes, have been successfully involved in the design of
tional guidance laws have been modified, yielding new guidance guidance and control systems [2330]. High-order robust exact
schemes such as augmented proportional navigation (APN), pure differentiators (HOREDs) have also been successfully employed as
proportional navigation, optimal guidance laws, predictive guidance observers in homing interceptor guidance loops. It was shown in [24]
laws, switched bias proportional navigation, and circular navigation that HOREDs give a good estimate of the target acceleration from
guidance [14]. Furthermore, new guidance strategies have been available measurements such as line-of-sight (LOS) range and angle,
evolved and implemented in navigation and homing guidance loops closing velocity, and interceptor orientation. The authors found that
using different control approaches such as nonlinear control, modern the guidance law using HOREDs results in less maneuverability ratio
close to the endgame of the homing loop as compared to the APN
Received 23 April 2013; revision received 31 December 2013; accepted for guidance law. In [31], it was also found that control loops including
publication 6 January 2014; published online 11 June 2014. Copyright HOREDs achieved better performance when compared to those
2013 by the American Institute of Aeronautics and Astronautics, Inc. All using EKF. It is worth noting that homogeneous or arbitrary-
rights reserved. Copies of this paper may be made for personal or internal use,
order sliding-mode controllers (AOSMCs) and HOREDs show
on condition that the copier pay the $10.00 per-copy fee to the Copyright
Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923; include impressive characteristics such as high tracking accuracy, finite-time
the code 1533-3884/14 and $10.00 in correspondence with the CCC. convergence, robustness against disturbances, and the ability to cope
*Assistant Professor, Department of Aeronautical Engineering, P.O. Box with situations involving discontinuous dynamics, which is suitable
80204; bkada@kau.edu.sa. Member AIAA. with regard to the interception missions [3237].
1999
2000 KADA

In this paper, a preliminary investigation is performed to design an Px  Lrf x; f x0  0
Qx  Lg Lr1
integrated robust arbitrary-order sliding-mode homing guidance
(AOSMHG) system to successfully intercept highly maneuvering are some bounded continuous functions on an open set X xo  X
targets with near-to-zero miss distance, less interceptor acceleration surrounding a given initial condition xo with
demand, and minimum control effort. To reach this objective, a
strategy that combines AOSMCs, HOREDs, nonlinear engagement
jPxj C; K m Qx K M (4)
geometry, and nonlinear interceptor dynamics in one guidance and
control system is adopted. To implement this strategy, advanced
guidance laws [i.e., APN and direct collision or hit-to-kill (HTK)] are
used as sliding manifolds. In addition, and based upon the arbitrary- K m , K M , and C are some positive constants; and Lf  and Lg 
order feature of AOSMCs and HOREDs, the relative degrees of the are the Lie derivatives.
controllers and observers are artificially increased to enhance the Assumption A4: Under a local diffeomorphic coordinate transfor-
performance of the AOSMHG system, remove chattering effect, and mation (i.e., invertible diffeomorphism) k  k1 (k  1; : : : ; r),
improve observation accuracy. For realistic scenarios, the target nominal system (1) is feedback equivalent to the following
acceleration is estimated from available measurements using arbi- controllable canonical form:
trary orders of HOREDs. To the authors knowledge, AOSMHG
system is the first to focus on missile homing guidance design using _  Ax  Bxu (5)
AOSMCs, HOREDs, and nonlinear interceptor model.
The outline of the remaining of the paper is as follows. Section II with
Downloaded by University of Toronto on November 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.G000001

describes the mathematical formulation of HSMCs and HOREDs.


In Sec. III, nonlinear kinematics and dynamics of the interception A   Lf L2f ::: Lr1 Lrf T x (6)
f
process are formulated. The design and mechanization of the
AOSMHG strategy are presented in Sec. IV. The performance and
robustness of AOSMHG laws are analyzed and discussed in Secs. V
and VI, respectively. Finally, Sec. VII summarizes the key results and B  0 0 ::: 0 f  x0 
Lg Lr1 T
(7)
concludes the paper.

The result is that the trajectories of system (1) are understood in the
Filippov sense, and they are infinitely extendible in time for any
II. High-Order Sliding Modes: Control control input u provided that this control is a Lebesgue-measurable
and Observation function and that new auxiliary system (5) has a locally bounded
A. Problem Statement right-hand side.
Consider the following class of smooth nonlinear dynamic
systems, B. Homogeneous High-Order Sliding-Mode Controllers
Theorem 1 [32,34,35]: Let assumptions A1A4 hold; bounded r-
x_  fx  gxu (1a)
sliding-modes feedback controllers ur of the form

yx  x (1b) ur  Gr1;r ; _ : : : ; r1  (8)

where x Rn is the state vector, f and g are uncertain smooth can be constructed in new constraint space (5) to drive the output x
function vectors, u R is an affine control input, y  R is an and its successive time derivatives k k  1; : : : ; r 1  toward
observed or measured output signal, and n is the system dimension. their zero level in finite time and in spite of internal and external
Typical uncertainties include modeling errors, unmodeled dynamics, disturbances.
actuator and sensor nonlinearities, delays, measurement noises, and The magnitude G is an adjustable positive gain. Feedback
bounded external disturbances. equation (8) is the basic equation of a class of high-order sliding-
Let system (1) be closed by some dynamical discontinuous mode controllers that are defined by recursive procedures. One of the
feedback; based upon the Filippov solution trajectories, the control most impressive families of these controllers is the following quasi-
problem considered here aims to design high-order sliding modes continuous homogeneous r-sliding controllers constructed based on
that are able to fulfill the constraint x  0 after a finite-time the homogeneity approach:
transient and retain it afterwards in spite of internal and external
disturbances. To do so, the following assumptions are considered: r1;r
Assumption A1: The output constraint x is supposed to be of the ur  Gr1;r  G (9a)
N r1;r
r-order differentiability class, where r defines the relative degree of
x with respect to the control input u. The degree r is supposed to be The functions r and N r are computed using the following recursive
constant and known. procedure
Assumption A2: The function x and its successive time for k  0:
derivatives k x up to r 1 are continuous functions of x and form
the following nonempty integral set: 
0;r 
(9b)
r  fx Xjx  _ x  : : :  r1 x  0g N 0;r  jj

integral set (2) for 1 k < r 1,

where X Rn denotes the operating space of system (1). 


k;r  k  k N rkrk1 k1;r N k1;r
Assumption A3: If assumptions A1 and A2 hold, x satisfies the k1;r
(9c)
following equality: N k;r  j k j  k N rkrk1
k1;r

r x  Px  Qxu (3) It is noted that the homogeneity features of dynamic systems provide
for the highest possible asymptotic accuracy for real-time control
where (i.e., discrete measurements with delays and measurement noises).
KADA 2001

C. High-Order Robust Exact Differentiators R_  V R  V m cos m   V t cos t   (13a)


For a given relative degree r > 1, the real-time implementation of
the controllers [Eq. (9)] requires exact calculation or direct _  V R  V m sin m   V t sin t  R (13b)
measurement of (r 1)-order time derivatives of the output signal
x. The following high-order differentiators provide a practical
way to compute the signals k (k  1; : : : ; r 1) provided that r V_ R  V 2 R  Am sin m   At sin t    V 2 R  Am;R  At;R
is bounded [36,37]: (13c)

8
>
> for k  0
>
>
>
> z_  v0 ; v0  0 L1r jz0 jr1r signz0   z1
< 0
for 0 < k < r 1
(10)
>
> z_k  vk ; vk  k L1rk jzk vk1 jrk1rk signzk vk1   zk1
>
>
>
> for k  r 1
:
z_r1  r1 L signzr1 vr2 
Downloaded by University of Toronto on November 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.G000001

where zk k and k are the differentiators (or observers) param- V_  V V R R Am cos m   At cos t  
eters, L is a known positive Lipschitz constant, and vk (k  0; : : : ;
r 2) denote the successive outputs of the differentiators. The  V V R R Am;  At; (13d)
controllers [Eq. (9)] together with the observers [Eq. (10)] produce r-
order real-time output-feedback controllers of the form The polar coordinates and R are given by

ur  Gr1;r z; z_ : : : ; zr1  (11)  artanht hm xt xm  (14)

The new controllers [Eq. (11)] possess the following properties:


q
Property P1: In the absence of measurement noises, the observer-
based controllers [Eq. (11)] provide for the finite-time convergence of R xt xm 2  ht hm 2 0 (15)
the output tracking variable x and its successive (r 1)-time
derivatives to their zero level. The subscripts m and t denote motion information of the missile and
Property P2: The controllers [Eq. (11)] feature high accuracy in target, whereas the subscripts R and signify radial and tangential
the presence of sampling noise and delays, and they ensure strong components, respectively.
robustness against disturbances.
Property P3: In case of x measurements corrupted by an B. Nonlinear Missile Interceptor Dynamics
unknown measurable noise of magnitude (i.e., a bounded The missile interceptor considered in this study is a hypothetical
Lebesgue-measurable function of time), the convergence to the tail-controlled missile where the nonlinear longitudinal motion (i.e.,
boundary layers j k j < k rkr (k  0; : : : ; r 1) is assured for pitch-plane motion) is governed by the following equations [28,37]:
some constants k > 0.
Property P4: The parameters G, k , and k have to be adjusted to
_  KZ M2 CD CZ ; M; c  sin  g sin m
M (16a)
reach the desired level of performance, and they can be adapted in real 0
vs
time. The value of the Lipschitz constant is subject to the following
condition [32]:
g
L C  GK M (12) _  KZ MCZ ; M; c  cos  cos m  q (16b)
vs M

Property P5: The performance of the closed-loop control can be


considerably improved, and the chattering effects can be efficiently g
_ m  K Z MCZ ; M; c  cos cos m (16c)
removed by artificially increasing the relative degree r (i.e., the order vs M
of the controllers [Eq. (11)]).

q_  K M M2 CM ; M; c  (16d)
III. Kinematics and Dynamics of the Interception
Process
x_ m  Mvs cos m (16e)
In the geometric-based intercept approach, the missile and target
dynamics are combined with engagement kinematics to produce the
homing-missile engagement rules. By pointing its velocity vector at
the future position of the target, the intercept would benefit from
decreasing the required lateral acceleration and the final intercept time.

A. Planar Engagement Kinematics


The planar engagement model is often considered during the
endgame phase of the interception to derive the engagement kinemat-
ics. Assuming the target and interceptor as point masses moving at
constant speeds, one possible engagement process is presented in
Fig. 1 for which the kinematics are expressed, in polar coordinate
system, as follows: Fig. 1 Planar interceptor-target engagement geometry.
2002 KADA

2 3
h_m  Mvs sin m (16f) V m cos m   V t cos t  
6 V m sin m   V t sin t  R 7
F 6
4
7
5 (24c)
where vs is the sound speed, and g is the gravitational acceleration. As V 2 R
the effect of the aerodynamic control c on the force coefficient is V V R R
negligible, the aerodynamic coefficients CZ and CM are given as
follows [38]:
  L  0 0 sin m  cos m   (24d)
M
CZ ; M  an 3  bn jj  cn 2 (17a)
3  
 0 0 sin t   cos t   (24e)
 
8M
CM ;M;c   am 3  bm jj  cm 7   em q  dm c
3
(17b)  Am (24f)

The coefficients KZ and KM are computed as follows:


v  At (24g)
KZ  0.7P0 Smvs (18a)
The vector   R V R V T R4 is the interception state
Downloaded by University of Toronto on November 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.G000001

vector. In this representation,  Am is referred to as the control input


K M  0.7P0 SdI y (18b) to the guidance problem, whereas v  At is considered as external
disturbance [8]. The vector  R4 is the vector of matched
where P0 is the static pressure, S is the surface area, and I y is the pitch and unmatched disturbances, and  R is the guidance constraint
moment of inertia. Numerical values of various coefficients and that should be designed properly to satisfy the requirements of
physical parameters are available in [28,37,38]. State model (16) successful interception. The control problem represented by Eq. (24)
represents the dynamics of a single-input/multi-output system where is to seek for a feedback control law  v;  for which the
interception processes will be successfully achieved. To do so, the
x  M m q xm hm T (19) following assumptions are made:
Assumption A5: F  and L might be uncertain due to
interceptor parameterization uncertainties or target location and
y  M q Am T (20) target speed estimations.
Assumption A6: and are supposed to be unknown but bounded
in their Euclidian norms (i.e., k k2 max R and kk2
u  c (21) max R ).
Assumption A7: The case  0 corresponds to the nominal form
The output vector y is defined by the variables available for of model (24), and  0 corresponds to the nonmaneuvering target.
measurement M and q, and the missile normal acceleration Am
computed in (g) from B. Missile Lethality Model
A target tracking and interception mission can be successfully
Am  K Z M2 CZ ; M (22) achieved if the condition V R  R_ < 0 or V  0 (i.e., _  0) is
guaranteed. In the present study, the interception mission is
C. Actuator Dynamics
considered successfully achieved if the following lethality condition
(LC) is satisfied:
Actuators are inherently nonlinear devices exhibiting highly 
nonlinear dynamics of unknown order. To simplify the task, and 1 if Rf Rl
without exceeding the hardware capabilities, we incorporate the LC  (25)
0 if Rf > Rl
actuator dynamics into the guidance loop as a function of the relative
degree r through the following state model: where Rf denotes the value of the range R at the time of interception,
 and Rl denotes the lethal or kill radius that defines the zero-intercept
a _ k  k1 1 k r 1 condition depending on the target size. This condition also guaran-
(23)
a _ r  f; rc  tees the lethality model in terms of miss distance [17]. The engage-
ment process is interrupted when the first condition in model (25) is
where Rr and a > 0 are the state vector and time constant of the fulfilled and the corresponding time tf is considered as the impact
actuator, respectively, whereas f; c  could be a locally bounded time or the final intercept time. The two measures Rf and tf are
Borel-measurable function (rc is the input of the actuator).

IV. Arbitrary-Order Sliding-Mode-Based Homing


Guidance: Design and Mechanization
A. Problem Formulation
A general state-space form corresponding to guidance problem (13)
can be formulated in the polar coordinate system R;  as follows:

_  F   L  v   (24a)

  0 (24b)

with Fig. 2 Missile and seeker orientations in an inertial reference [39].


KADA 2003

Fig. 3 Block diagram for the real-time implementation of AOSMHG laws


p
considered among the indices that will be used in Secs. V and VI to DHK x  V  R;
_ LHK x  V c0 R (27)
evaluate the performance of the proposed guidance laws.
where c0 R .
Using the tracking functions APN , DHK , and LHK , three AOSM-
Downloaded by University of Toronto on November 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.G000001

C. Interception Sliding Manifold Design


based guidance laws, i.e., AOSM-APN, AOSM-DHK, and AOSM-
As the design of the sliding manifold x is a critical aspect of any LHK are constructed and implemented.
sliding-mode guidance and control law, the sliding manifold
candidates are selected herein as functions of the LOS rate . _ The
D. Homing Loop Elements
selection of _ as commanded output has a great importance for both
physical principle and mechanization of the guidance laws, as is In missile dynamic model (16), the normal acceleration Am has a
shown in the Sec. IV.E. relative degree r  2 with respect to the fin deflection c [28]. As the
relative degree of the outputs V and V R with respect to Am is 1 [see
1. APN-Based Guidance interception model (24)], the result is that the inputoutput dynamics
V c and V R c both have a relative degree r  3. This degree
It is well known that APN guidance attempts to minimize the miss
is considered as the minimum order required for the construction of the
distance via mechanization of the missile acceleration command Am
controllers [Eq. (11)] that could be artificially increased for enhanced
normal to its velocity vector. By adding a term to compensate for
performance and robustness. In the sequel, third- and fourth-order
target acceleration, APN guidance law yields superior performance
controllers, differentiators (observers), and actuators are constructed to
against a highly maneuvering target when compared to the traditional
compute the commanded fin deflection c to achieve a successful
navigation laws. For this reason, an APN-based sliding manifold
interception:
candidate is taken as
Controllers:
 A
APN x  N V R _  t; Am c  Gr1;r z0 ; z1 ; : : : ; zr1  (28)
2
 A
t
 N V R _  cos t   Am (26) with
2

where At;  At cos t   denotes the target acceleration trans- c  G3  G N2;3
2;3
for r  3
versal to the LOS. To ensure good dynamic performance, the (29)
c  G4  G N3;4 for r  4
navigation constant N is generally selected in the range of 3 to 5. 3;4

8
>
> 0;r  z0  ; N 0;r  jz0 j  jj
< for k  1; :::;r1
rkrk1 (30)
>
> k;r  zk k N k1;r k1;r N k1;r ; N k;r  jzk jk N rkrk1
k1;r
: k
 zk
Differentiators (observers):
8
>
> for k  0; :::;r2
<
z_k  vk ;vk  k L1rk jzk vk1 jrk1r1 signzk vk1 zk1 ; v1 
(31)
> z_2  2 Lsignz2 v1 
> for r  3
:
z_3  3 Lsignz3 v2  for r  4

Table 1 Simulation parameters for an


Evasive Circular Maneuver
2. Hit-to-Kill-Based Guidance
To reduce the weight and cost of todays interceptor missiles, the Interceptor Target
HTK guidance is proposed to kill the threat directly by body-to-body V m0  3M V t  2M
contact without need of the warhead. In terms of sliding-mode Am0  0g At  10g
manifold candidates, two velocity tracking errors are used to imple- m0  86 deg t0  0 deg
xm0  0 m xt0  0 m
ment direct hit-to-kill (DHK) guidance and less-aggressive hit-to-kill
hm0  3000 m ht0  10; 000 m
(LHK) guidance, respectively [24]:
2004 KADA
Downloaded by University of Toronto on November 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.G000001

Fig. 4 Circular air-to-air interception: a) engagement trajectories, b) LOS range R, c) acceleration demand, and d) control effort.

Linear actuators: loop autopilot is used to improve the transient response of the missile
interceptor. From state model (16) and aerodynamic model (17b), the
8
> _  2 inputoutput dynamics c qc are derived as follows:
>
>.a 1
< 2
.. _
qK M M CM ;M;c 
(32)    
>
> _  r 8M
>
: a _ r1
a r  a0 1 : : : ar1 r  rc KM M2 am 3 bm jjcm 7 em qK M M2 dm c
3
(33)
The function in Eqs. (30) and (31) denotes indifferently the sliding-
mode variables APN , DHK , or LHK. The conditions [Eq. (4)] and the As the relative degree of the inner-loop output is rq  1, a first-order
Lipschitz constant [Eq. (12)] associated to the sliding variables sliding-mode controller could be designed using the pitch rate
[Eq. (26)] and [Eq. (27)] are illustrated in Appendix A. tracking error q  q qc as a sliding manifold. Regulation
condition (3) becomes
E. Inner-Loop Autopilot Design
The stability of the outer-loop sliding-mode controllers requires an _ q  q_ q_ c
     
inner-loop controller design to follow the pitch rate commands qc 8M
produced in these outer-loop controllers. As the major aim of  KM M2 am 3  bm jj  cm 7   em q q_ c
3
interception laws (26) and (27) is to follow the acceleration or speed
commands, which are realized in the outer-loop control, the inner-  K M M 2 dm c (34)

Table 2 Homing performance for circulation air-to-air interception


Guidance law Interception time tf , s Final range Rf , m Miss distance jht hm j, m Maxjc j, deg MaxjAm j, g V m range, M
AOSM-APN third order 20.723 0.410 0.358 2.635 8.4687 [3.000, 3.035]
AOSM-APN fourth order 20.755 0.117 0.106 2.667 8.4301 [2.996, 3.030]
AOSM-DHK third order 20.832 0.189 0.109 2.755 8.4669 [2.991, 3.020]
AOSM-DHK fourth order 20.702 0.004 0.004 2.607 8.4167 [3.000, 3.038]
AOSM-LHK third order 20.827 0.121 0.064 2.749 8.4642 [2.991, 3.021]
AOSM-LHK fourth order 20.702 0.030 0.029 2.607 8.4168 [3.000, 3.038]
KADA 2005

with KM M2 dm 0. Using the controller topology [Eq. (9)], the Table 3 Simulation parameters for
inner-loop autopilot has the following form: interception of a target with intelligent evasion
capabilities
Interceptor Target
c;q
r1  G sign 
q (35)
V m0  1.5M V t  1.5M
Am0  0g At0  0g
m0  0 deg t0  180 deg
As the controllers [Eq. (9)] establish finite-time stable r-sliding-mode xm0  0 m xt0  2500 m
 0 [31,33], the pitch rate dynamics _ q convergence is zero level in hm0  0 m ht0  0 m
finite time.
Downloaded by University of Toronto on November 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.G000001

Fig. 5 Time histories of sliding variables and their time derivatives for third- and fourth-order controllers: a) APN sliding modes, b) direct-hit-to-kill
sliding modes, and c) less-hit-to-kill sliding modes.
2006 KADA
Downloaded by University of Toronto on November 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.G000001

Fig. 6 Evasive interception: a) engagement trajectories, b) LOS range R, c) acceleration demand, and d) control effort.

F. Mechanization of AOSMHG Laws and Conceptual Solutions body centerline, respectively. Figure 2 illustrates the different angles
Central to the mechanization of any intercepting guidance law is needed to the implementation of such an approach.
the construction of the LOS angular rate .
_ Instead of an analytical In [40], the authors found that using models (36)(37) to
derivation of expression (14), which requires reconstruction of reconstruct _ yields an excessive amplification of measurement noise
unmeasured parameters, a numeric derivative of could be obtained if the responsiveness of the interceptor to the commands is slow. As
by passing it through the following derivative filter: an alternative, in the present study, the signal _ is constructed from
Eqs. (14) and (36), which need only the measurement of missile and
s
_ s target positions. In the absence of an accurate model, xM , hM , V M ,
 (36) and m could be provided by an inertial measurement unit (IMU) and
s f s  1
GPS system, and the relative degree r could be estimated from input
where f is the time constant of the filter, and s is the Laplace
operator. This numerical derivation yields better performance than Table 5 Simulation parameters for
the analytical one [39]. The LOS angle in Eq. (36) could be interception of weaving target
constructed from seeker measurements as follows [40]:
Interceptor Target
Z Z
V m0  1.7M V t  1M
    _ dt;      _ dt (37) Am0  0g t max  11.5 deg
m0  15 deg  4.5 rads
xm0  0 m xt0  10; 000 m
where the overbar denotes a measured quantity; and , , and are hm0  0 m ht0  5000 m
the seeker dish angle, gimbal angle, and inertial angle to the missile

Table 4 Homing performances for interception of a target with intelligent evasion capabilities
Guidance law Interception time tf , s Final range Rf , m Miss distance jht hm j, m Maxjc j, deg MaxjAm j, g V m range, M
AOSM-APN third order 2.772 0.252 0.013 14.892 6.743 [1.496, 1.503]
AOSM-APN fourth order 2.770 0.243 0.005 13.515 6.285 [1.498, 1.504]
AOSM-DHK third order 2.767 0.277 0.201 13.975 6.321 [1.500, 1.517]
AOSM-DHK fourth order 2.767 0.141 0.007 11.655 5.659 [1.500, 1.501]
AOSM-LHK third order 2.767 0.171 0.125 13.460 6.508 [1.499, 1.511]
AOSM-LHK fourth order 2.767 0.049 0.014 11.600 5.636 [1.500, 1.510]
KADA 2007
Downloaded by University of Toronto on November 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.G000001

Fig. 7 Weaving interception: a) engagement trajectories, b) LOS range R, c) acceleration demand and d) control effort.

output measurements. The differentiator [Eq. (36)] is used as a interception model (24) (i.e.,  0). For comparison purposes, the
guidance filter to compute the LOS rate _ out of the flight control AOSMHG laws are applied simultaneously for each scenario using
system and independent from the interceptor dynamics, as shown two relative degrees (i.e., r  3 and r  4).
in Fig. 3
A. Target Performing an Evasive Circulation Maneuver
In this scenario, the missile is supposed to intercept a target that
V. Homing Performance Analysis performs an evasive circular trajectory governed by the following
To give perspective to the proposed guidance strategy and to equations of motion:
understand how interception key elements affect the accuracy and
robustness of the AOSMHG loop, we carry out, in this and the next x_ t  V t cost t  t0  (38a)
section, many computer simulations involving 1) a target with known,
evasive, and agile maneuvers; 2) interceptor dynamics with large error
modeling and parameter uncertainties; and 3) target maneuver estima- h_t  V t sint t  t0  (38b)
tion from noise-corrupted measurements,. The intent of simulating
various interception scenarios is to measure the accuracy, agility,
maneuverability, and adaptability of each AOSMHG law for various _ t  t  At V t (38c)
classes of problems. In all simulations, the lethality radius Rl in
Eq. (25), the guidance constant N in Eq. (26), and the constant c0 in The engagement process starts from the initial conditions given in
Eq. (27) are set to be Rl  0.5 m, N  3, and c0  1, respectively. Table 1. The engagement trajectories, LOS range, interceptor acceler-
In this section, the three selected scenarios are run assuming ation demand, and control efforts are shown, respectively, in Fig. 4.
perfect knowledge of target maneuvers and using the nominal form of Detailed performance indices are presented in Table 2.

Table 6 Homing performances for interception of weaving target


Guidance law Interception time tf , s Final range Rf , m Miss distance jht hm j, m Maxjc j, deg MaxjAm j, g V m range, M
AOSM-APN third order 44.042 0.268 0.219 4.582 4.439 [1.701, 1.999]
AOSM-APN fourth order 41.951 0.217 0.110 3.636 3.614 [1.701, 2.060]
AOSM-DHK third order 42.099 0.322 0.071 5.899 3.249 [1.694, 2.147]
AOSM-DHK fourth order 42.338 0.166 0.020 3.515 2.936 [1.701, 2.065]
AOSM-LHK third order 41.534 0.218 0.188 6.713 3.630 [1.687, 2.150]
AOSM-LHK fourth order 41.098 0.199 0.024 3.376 2.883 [1.701, 2.092]
2008 KADA

Table 7 Simulation parameters for Table 8 Homing performance comparison


homing synthesis
Miss
Interceptor Target Interception distance Maxjc j, MaxjAm j,
V m0  1.21M V t  1.21M Guidance law time tf , s jht hm j, m deg g
Am0  0g At max  20g AOSM-APN 1.444 0.178 16.121 3.910
m0  25 deg t  0.05 s AOSM-DHK 1.444 0.027 16.303 3.950
xm0  0 m xt0  1000 m AOSM-LHK 1.444 0.062 16.364 3.963
hm0  0 m ht0  0 m Zeros-effort miss 1.490 0.300 30.000 55.00

The time histories of the sliding variables [Eqs. (26) and (27)] and C. Target with Sinusoidal Change in Heading (Weaving Target)
their successive time derivatives up to (r 1) are depicted in Fig. 5. In this case, and from the initial conditions shown in Table 5, the
target is supposed to perform a sinusoidal heading maneuver of the form
B. Target with Intelligent Evasion Capabilities
In this scenario, the target is supposed to know the position of the _ t  t;max sint (40a)
pursuer, and it would perform jumps or abrupt changes in the
direction of motion as follows:
At  _ t V t (40b)
if R > R1 : h_t  0 (39a)
Downloaded by University of Toronto on November 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.G000001

if R1 < R R2 : h_t 0 (39b) Figure 7 shows close-up views of trajectories, LOS range,
interceptor normal acceleration, and control effort; and Table 6 gives
detailed results of the homing performance.
if R R1 : h_t  0 (39c) According to the preceding results, the proposed AOSMHG
laws demonstrate good performance and successfully fulfill lethality
Both the interceptor and target are supposed to fly straight courses, model (24). On the other hand, it appears that artificially increasing
starting from the initial conditions given in Table 3 with R1  500 m the relative degree of the controllers [Eq. (11)] improves homing
and R2  375 m. Figure 6 depicts the time histories of the engage- performance indices mainly in the interception time tf , acceleration
ment parameters, and Table 4 presents the performance indices. demand Am , and control effort c , and it removes the chattering effects.

Fig. 8 Sample run for homing synthesis: a) engagement trajectories, b) LOS range, c) acceleration demand, and d) control effort.
KADA 2009
Downloaded by University of Toronto on November 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.G000001

Fig. 9 AOSMHG laws sensitivity to parameter uncertainty: a) flight time, b) miss range, c) control effort, and d) acceleration demand.

VI. Performance Comparison and Robustness Issues measurement noises are discussed. Based upon the results obtained in
The engagement scenarios illustrated in the previous section were the previous section, only the fourth-order AOSMHG laws are
accomplished without including disturbances, target maneuver applied in the following scenarios.
estimation, and measurement noises. In the present section, compar-
ison to other guidance systems, sensitivity to parametric uncertainties A. Performance Comparison and Sensitivity to Modeling Errors
and modeling errors, real-time observation, and robustness against To compare the performance of the proposed AOSMHG laws with
other guidance systems, a scenario similar to the one proposed in [41]
is considered. The simulation parameters are selected as shown in
Table 9 Simulation parameters for Table 7. The engagement trajectories are plotted in Fig. 8, and a
interception using target maneuver performance comparison is shown in Table 8.
estimation
To evaluate the sensitivity of the AOSMHG laws to parameter
Interceptor Target uncertainties and modeling errors, a series of simulations that
V m0 : 3M V t : 3M correspond to different values of the bound max of the uncertainties
Am0 : 0g At0 : 0g  in Eq. (24) are carried out and a polynomial fitting curve is used.
m0 : 45 deg t0 : 180 deg The results of these simulations are shown in Fig. 9.
xm0 : 0 m xt0 : 25,000 m The results shown in Table 8 indicate that AOSMHG laws
hm0 : 0 m ht0 : 20,000 m guarantee better interception performance as compared to the guid-
ance system proposed in [41]. From Fig. 9, it is apparent that

Table 10 Homing performance of interception with target maneuver estimation


Guidance law Interception time tf , s Final range Rf , m Miss distance jht hm j, m Am;max At;max
AOSM-APN without estimation 20.7260 0.3375 0.3108 0.91378
AOSM-APN with estimation and without noise 20.7250 0.2198 0.1370 0.83975
AOSM-APN with estimation and with noise 20.7400 0.2077 0.0458 1.68787
AOSM-DHK without estimation 20.7260 0.1484 0.1216 0.84181
AOSM-DHK with estimation and without noise 20.7250 0.1405 0.1389 0.83975
AOSM-DHK with estimation and with noise 20.7400 0.1728 0.0045 1.68723
AOSM-LHK without estimation 20.7260 0.1579 0.1572 0.84121
AOSM-LHK with estimation and without noise 20.7250 0.0527 0.0465 0.83975
AOSM-LHK with estimation and with noise 20.7400 0.0578 0.0535 1.68737
2010 KADA

Fig. 10 Target A^ t estimation using HOREDs.


Downloaded by University of Toronto on November 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.G000001

Fig. 11 Interception with target estimation and measurement noises.

parameter uncertainties and modeling errors have a nonlinear effect framework for the design of real-time enhanced homing-missile
on the AOSMHG performance. It follows that, even if the sliding guidance laws. Using the arbitrary-order feature of these controllers
manifolds [Eqs. (26) and (27)] are sensitive to the unknown bounded and observers, both APN-based and HTK-based AOSMHG schemes
uncertainties , the AOSMHG laws compensate these uncertainties have shown their potential to successfully intercept highly maneu-
very well. verable targets during endgame scenarios. Simulation results proved
that the AOSMHG yields high collision accuracy and substan-
B. Sensitivity to Target Maneuver Estimation and Measurement Noises tial improvement of homing-missile guidance performance, such as
To simulate realistic interception missions and evaluate the near-to-zero miss-distance, reduced maneuverability ratio require-
sensitivity of the AOSMHG to measurement noises, a target that, ments, and unsaturated control effort, even when considering the
being aware of an interception attempt, randomly performs an unknown shape of the target maneuver, target maneuver estimation, or
evasive maneuver to escape from the pursuer is considered. The target measurement noises. In addition, the AOSMHG has demonstrated
executes an evasive maneuver of 5g-turn down followed by 10g- high robustness against parametric uncertainties. It was also found that
weave change in the normal acceleration. artificially increasing the relative degree of the AOSMHG laws results
In this simulation, the target acceleration At; is constructed from its in enhanced control performance, high observation accuracy, and
position and orientation measurements using the observers [Eq. (10)], smooth control signal. Finally, the following topics will be considered
and the time derivative _ is computed using the derivative filter for future works: three-dimensional implementation with more evasive
[Eq. (36)] with time constant f  0.001 s. The scenario parameters scenarios, multiple target interception, Monte Carlo runs, and more
are defined in Table 9. investigation of the missile interceptor dynamic model.
An angular zero-mean additive white Gaussian-distributed noise
with a standard deviation of 1 deg is added to the measured _ in the
seeker system. Appendix:
As can be seen from Fig. 10, the accuracy and convergence of the
target maneuver estimate increase with an increase in the order of the
observers [Eq. (10)]. Figure 11 shows that interception trajec- A1 Sliding Mode Variable DHK  R1
tories obtained by the three AOSMHG laws are similar, and from
Table 10 it can be seen that AOSMHG with target maneuver The successive time derivatives up to r order of the expression
estimation yields minimum miss distance and, in the absence of R1 are given by the following recursive procedure
measurement noises, results in less missile acceleration demand as 8
compared to the one without estimation of target acceleration. >
> R1 1  R1 1  R2
>
< R1 2  R2 1  2R1 2  R3
.. (A1)
>
> .P
>
:
VII. Conclusions R1 r  Rr 1  r 1kr1 Rk rk1  Rr1
This paper demonstrates that homogeneous high-order sliding-
mode controllers and observers can provide a convenient and effective In case of r  3
KADA 2011

3 1 3


DHK  R   R3 1  3R2 2  3R1 3  R4 2V V 3R 2V V 2R 2V R
PLHK  A2
t;   A  2A1 1
m sin m m
(A2) R3 R2 R t;
3c0 V R V 2 3c V 3R c V
Using the interception model (13), the expression A2 becomes A2
m cos m  p  0 2 p   0p A
4 R R 2 8 R R R R t;
 1  2  3 3c V c c0 1
V V V V 0 pR At;R  p0 A1
t;R  p Am sin m (A11)
3 2
DHK  V R  3V 1
R  3V R R 4 R R 2 R 2 R
R R R R
2V V 3R 2V V 2R 2V R
  Am;  At;  A2 2
m;  At; (A3)
R3 R2 R 2V R
QLHK  cos m  cos m 1 2 2
m   sin m m
R
With m  m , the second time derivative of the missile V 3 VR 1
2
acceleration transversal to the LOS Am; is p cos m p sin m  p cos m _ m (A12)
R R 4 R R 2 R
A2 2 1 1
m;  Am cos m 2Am sin m m The inequalities [Eq. (4)] and the Lipchitz constant in [Eq. 12] are
Am cos m 1 2 2
m   sin m m  (A4) approximated by

2V max max 3
V R  2V max V max 2 2V max
and the expression (A3) jPLHK j A2;max
t;   2R  R Amaxt;
Downloaded by University of Toronto on November 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.G000001

R3l Rl Rl
2V V 3R 2V V 2R 2V R 3c0 V max max 2 3c V max 3
3 A A2 R V 
2
DHK   t; Am cos m  2A1
m m
1;max
 A2
m  p  0 2Rp
R3 R2 R t; 4 2
Rl Rl 8 Rl Rl
 
2V R
2A1 1
m sin m m  cos m cos m 1 2 2
m  sin m m Am c0 V max 3c V max
c
R  p Amax 0 p R 1;max
 Amax  p0 At;R
Rl Rl t; 4 Rl Rl t;R 2 Rl
 PDHK QDHK Am (A5) c cK
 0p2Z A1
m  CLHK (A13)
2 Rl
Under the constraints jAk k max
m; j Am , jAk k;max
t; j At; , Rl R
Rt  0  R0 0, j cos j 1, j sin j 1, and k;min k
min

min

k;max , the functions PDHK and QDHK in [Eq. A5] are bounded
2V

V

3V min

K m;LHK 

 j1;min m 2  2;min
m j

p 

R
p 

functions and verify the inequalities [Eq. (4)] R0 R0 R0 4R0 R0





max








1

2V


2V V 3R

2V V 2R

2V R


p _ min
QLHK
R
 j1;max 2  2;max j
jPDHK j jA2 j




A
 jA2
m cos m j 2 R0 m

R
m m
t;
R3

R2

R t
l

max




V

3V max

1

2V max max 3
V R  2V max V max 2 

p 

R
p 

p _ max

m
 K M;LHK (A14)
 2jA1 1 2;max
m sin m m j At;  3
 2R Rl Rl 4Rl Rl 2 Rl
Rl Rl
2V max
 R
Amax
t  A2;max
m  2A2;max
m 1;max
m  CDHK (A6)
Rl 2V max max 3
V R 
LDHK A2;max
t;  3
 A2;max
m  2A1;max
m 1;max
m
Rl
2V min V max
3V max
K m;DHK  R
 1;min 2  2;min QDHK  1;max
m 2  2;max
m  p   pR
 (A15)
R0
m m Rl Rl 4Rl Rl
2V max
R
 1;max
m 2  2;max
m  K M;DHK (A7)
Rl

With 0 < Rl 1 and sup jr1;r j  1, the Lipchitz constant in A3 Sliding Mode Variable APN 
A
[Eq. 12] is approximated by NV R 1  2t;  Am
The third time derivative of the sliding mode variable APN is
2V max V max 3
LDHK A2;max
t;  3R  A2;max
m     1  2
Rl 3 V 2 V 1 V
3
APN  N V R  3V R  3V R
R R R
 2A1;max
m 1;max
m  1;max
m 2  2;max
m (A8)  3   
V N 3 V
 VR  At; A3 m  NV R
3
R 2 R
p
A2 Sliding Mode Variable LHK  R1 c0 R NV 2
3 R
RV 1 VRV
For the sliding mode variable LHK , [Eq. (3)] is written as follows R2
 
  NV 1 2 1 1 2V 2R V
p 3 3 2 R
RV V R V 2V R V 
3
LHK  R 1 3 c
0 R  PLHK  QLHK Am (A9) R R

NV R 3 V V R  3V R V  3V R V 2 
2 1 1
V
With the following time derivative R R
1 1 
   2  2V R V V R 2V 2R V 4V 3R V  2V 3R V
p 3 V 3 V R V 1 3 V 3R 
c0 R  c0 pR R
p  p 
 (A10) R2 R3
2 R 4 R R 8 R2 R
N
 A3 A3
m  PAPN  QAPN Am A16
One can find that 2 t;
2012 KADA

With Acknowledgments
This work was funded by the Deanship of Scientific Research
A2 2 1 1
m;R  Am sin m  2Am cos m m (DSR), King Abdulaziz University, Jeddah, under grant number 135-
030-D1434. The author, therefore, acknowledges with thanks the
Am sin m 1 2 2
m   cos m m  (A17)
DSR for technical and financial support.
One can find that
  References
V 2
V 1
R   At;R  sin m Am  P1  Q1 Am (A18) [1] Nesline, F. W., and Zarchan, P., A New Look at Classical vs. Modern
R
Homing Missile Guidance, Journal of Guidance, Control, and
Dynamics, Vol. 4, No. 1, 1981, pp. 7885.
  doi:10.2514/3.56054
3V R V 2 2V [2] Lin, C. F., Modern Navigation, Guidance and Control Processing,
V 2   1 1
A  Am cos m  At;R PrenticeHall, Englewood Cliffs, NJ, 1991, pp. 252413.
R
R2 R t; [3] Ghose, D., On the Generalization of True Proportional Navigation,
 
2V IEEE Transaction on Aerospace and Electronic Systems, Vol. 30, No. 2,
 cos m sin m _ m Am  P2  Q2 Am (A19) 1994, pp. 545555.
R
Downloaded by University of Toronto on November 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.G000001

2  3
3V 4 12V 2R V 2
R3
  4VRR2V At; R2 A2m;  A2t;   2VR A1
R3
6V R V
R3
1
m cos m  At; 
V 3
R 
4 5
3V 2 1 1 2 1 1
R3 Am sin m  At;R   Am sin m  2Am cos m m
 
4 2V 3V 2
 cos m At;  sin m _ m 3 cos m _ m sin m 1 2 2
m  cos m m Am  P3  Q3 Am (A20)
R R R

 
VRV
V 1
  At; Am cos m  P4  Q4 Am (A21)
R

2 3
2V V 2R V 3  
V V VR V
V 2
4 R
A A A1
m cos  A 1 5  cos sin  sin
_
m m Am  P5  Q5 Am (A22)
R2 R t; R t;R m t;
R m
R m

2  3
4V V 3R 2V 3 V R  2 
V V3 V V3 VR V 2 V V3
6 2 R 3 R   3 R2 2 R2 At; 
4 RR 3 3 RR 3 4V
R  2 R2 At;R R At; At;R
R V V R V 2 7
V 3
4
6 R3 7
5
 
1 1 1 1 2 1 1 2
 R Am; Am;R R Am cos m  At; R Am sin m  At;R  Am cos m 2Am sin m m  At;
2 VR V

2   3
V 2R V 2
6 3 R 2 2 R 2 cos m  4V R V
R  2 VR V
R 2 sin m 2
R A t;R cos m  A t; sin 
m 7
4 5Am  P6  Q6 Am (A23)
1 2 2
R sin m m  R cos m m cos m m  sin m m
VR _ V _

After certain rearrangements, the functions PAPN and QAPN in [Eq. A16] take the forms
2 3
R3 P2 P4  Q2 Q4 A2m 2 R3 P1 P5  Q1 Q5 A2m   3V
R2

P21  Q21 A2m 
6 9V R 8V 2R V 7
4  R2 P1 P4  Q1 Q4 Am  R3 P1  R2 P2 R P3
4V R V V
PAPN  N 6 2 7
5 (A24)
2V 3 3V 2R 4V 4R V 2V 4R V A3 A3
 R3R P4  R2
P5 VR
R P6  R3
 R4
 t;
2 m
N

2 3
R3 P2 Q4  P4 Q2  R3 P1 Q5  P5 Q1   R92 V R P1 Q4  P4 Q1   R22 P1 Q1 V
QAPN  N 4 8V 2R V 2V 3R 3V 2R
5 (A25)
R3
Q1  4VRR2V Q2 VR Q3  R3
Q4  R2
Q5 VRR Q6

The constraints jPAPN j CAPN and Km;APN QAPN K M;APN in [Eq. (4)] could be easily verified using the bounds of the functions Pi and Qi
(i  1; : : : ; 6). The Lipchitz constant satisfies
2 A3;max 3
3;max 2V 4;max V max 8V 2;max V max 2V 3;max
t;
2  Am  R
R4l

 R
R3l

Pmax
1  R
R3
Pmax
4
LAPN N 4 5 (A26)
8V 2;max V max 2V R3;max 4;max max
4V R V
 R
R3l

Qmax
1  R3l
Qmax
4  R3l
KADA 2013

[4] Zarchan, P., Tactical, and Strategic Missile Guidance, Progress in Guidance, Control, and Dynamics, Vol. 33, No. 3, 2010, pp. 695706.
Astronautics and Aeronautics, 3rd ed., Vol. 176, AIAA, Reston, VA, 1997, doi:10.2514/1.47276
p. 599. [23] Shtessel, Y. B., Shkolnikov, I. A., and Levant, A., Smooth Second-
[5] Cloutier, J. R., Evers, J. H., and Feeley, J. J., Assessment of Air-to- Order Sliding Modes: Missile Guidance Application, Automatica,
Air Missile Guidance and Control Technology, Control Systems Vol. 43, No. 8, 2007, pp. 14701476.
Magazine, Vol. 9, No. 6, 1989, pp. 2734. doi:10.1016/j.automatica.2007.01.008
doi:10.1109/37.41440 [24] Tournes, C. H., and Hanks, G., Hypersonic Glider Control Using
[6] Inncenti, M., Nonlinear Guidance Techniques for Agile Missiles, Higher Order Sliding Mode Control, IEEE Southeastcon, IEEE,
Control Engineering Practice, Vol. 9, No. 10, 2001, pp. 11311144. Piscataway, NJ, April 2008, pp. 274279.
doi:10.1016/S0967-0661(01)00094-6 [25] Foreman, D. C., Tournes, C. H., and Shtessel, Y. B., Integrated Missile
[7] Vathsal, S., and Sarkar, A. K., Current Trends in Tactical Missile Flight Control Using Quaternions and Third-Order Sliding Mode
Guidance, Defence Science Journal, Vol. 55, No. 2, 2005, pp. 265280. Control, 11th International Workshop on Variable Structure Systems
[8] Moon, J., and Kim, Y., Design of Missile Guidance Law via Variable (VSS), IEEE Conference Publications, June 2010, pp. 370375.
Structure Control, Journal of Guidance, Control, and Dynamics, [26] Scott, J. E., and Shtessel, Y. B., Launch Vehicle Attitude Control Using
Vol. 24, No. 4, 2001, pp. 659664. Higher Order Sliding Modes, AIAA Guidance, Navigation, and
doi:10.2514/2.4792 Control Conference, AIAA Paper 2010-7724, Aug. 2010.
[9] Shashi, R. K., Sachit, R., and Debasish, G., Sliding-Mode Guidance [27] Kada, B., Outer-Loop Sliding Mode Control Approach to Longitudinal
and Control for All-Aspect Interceptors with Terminal Angle Con- Autopilot Missile Design, Preprints of the 18th IFAC World Congress,
straints, Journal of Guidance, Control, and Dynamics, Vol. 35, No. 4, Vol. 18, Aug.Sept. 2011, pp. 1115711164.
2012, pp. 12301246. doi:10.3182/20110828-6-IT-1002.03818
doi:10.2514/1.55242 [28] Kada, B., Higher Order Sliding Mode Control for Missile Autopilot
[10] Gholson, N. H., and Moose, R. L., Maneuvering Target Tracking Using Design, World Academy of Science, Engineering and Technology,
Downloaded by University of Toronto on November 19, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.G000001

Adaptive State Estimation, IEEE Transactions on Aerospace and No. 70, Oct. 2012, pp. 174178.
Electronic Systems, Vol. AES-13, No. 3, 1977, pp. 310317. [29] Yamasaki, T., Balakrishnan, S. N., and Takano, H., Integrated Guidance
doi:10.1109/TAES.1977.308399 and Autopilot for a Path-Following UAV via High-Order Sliding Modes,
[11] Vergez, P. L., and Liefer, R. K., Target Acceleration Modeling for American Control Conference, IEEE, June 2012, pp. 143148.
Tactical Missile Guidance, Journal of Guidance, Control, and [30] Shkolnikov, I. A., Shtessel, Y. B., Zarchan, P., and Lianos, D. P.,
Dynamics, Vol. 7, No. 3, 1984, pp. 315321. Simulation Study of the Homing Interceptor Guidance Loop with
doi:10.2514/3.19861 Sliding Mode Observers Versus Kalman Filter, Proceedings of AIAA
[12] Zarchan, P., Representation of Realistic Evasive Maneuvers by the Guidance, Navigation, and Control Conference, AIAA Paper 2001-
Use of Shaping Filters, Journal of Guidance, Control, and Dynamics, 4216, Aug. 2001.
Vol. 2, No. 4, 1979, pp. 290295. [31] Levant, A., Homogeneity Approach to High-Order Sliding Mode
doi:10.2514/3.55877 Design, Automatica, Vol. 41, No. 5, 2005, pp. 823830.
[13] Uhrmeister, B., Kalman Filter for a Missile with Radar and/or Imaging doi:10.1016/j.automatica.2004.11.029
Sensor, Journal of Guidance, Control, and Dynamics, Vol. 17, No. 6, [32] Shtessel, Y. B., Fridman, L., and Zinober, A., Higher Order Sliding
1994, pp. 13391344. Modes: Editorial, International Journal of Robust and Nonlinear
doi:10.2514/3.21353 Control, Special Issue on Advances in Higher Order Sliding Mode
[14] Pan, S., Su, H., Chu, J., and Wang, H., Applying a Novel Extended Control, Vol. 18, Nos. 45, 2008, pp. 381384.
Kalman Filter to MissileTarget Interception with APN Guidance Law: [33] Levant, A., and Pavlov, Y., Generalized Homogeneous Quasi-
A Benchmark Case Study, Control Engineering Practice, Vol. 18, Continuous Controllers, International Journal of Robust and Nonlinear
No. 2, 2010, pp. 159167. Control, Vol. 18, Nos. 45, 2008, pp. 385398.
doi:10.1016/j.conengprac.2009.09.010 doi:10.1002/rnc.1199/(ISSN)1099-1239
[15] Zhao, H., Liu, W. J., Yang, J. H., and Xuan, Y. B., Design of Stochastic [34] Levant, A., Robust Exact Differentiation via Sliding Mode Technique,
Sliding Mode Variable Structure Guidance Law Based on Adaptive Automatica, Vol. 34, No. 3, 1998, pp. 379384.
EKF, Procedia Engineering, Vol. 23, Dec. 2011, pp. 125130. doi:10.1016/S0005-1098(97)00209-4
doi:10.1016/j.proeng.2011.11.2477 [35] Levant, A., Arbitrary-Order Sliding Modes with Finite Time
[16] Oshman, Y., Shinar, J., and Avrashi-Weizman, S., Using a Multiple Convergence, Proceeding of the 6th IEEE Mediterranean Conference
Model Adaptive Estimator in a Random Evasion Missile/Aircraft on Control and Systems, IEEE, Alghero, Sardinia, Italy, June 1998,
Encounter, Journal of Guidance, Control, and Dynamics, Vol. 24, pp. 349354.
No. 6, 2001, pp. 11761186. [36] Levant, A., Higher Order Sliding Modes, and Arbitrary-Order Exact
doi:10.2514/2.4833 Robust Differentiation, Proceedings of the 6th European Control
[17] Shima, T., Oshman, Y., and Shinar, J., Efficient Multiple Model Adaptive Conference, Portugal, 2001, pp. 9961001.
Estimation in Ballistic Missile Interception Scenarios, Journal of [37] Xin, M., and Balakrishnan, S. N., Nonlinear H Missile Longitudinal
Guidance, Control, and Dynamics, Vol. 25, No. 4, 2002, pp. 667675. Autopilot Design with -D Method, IEEE Transactions on Aerospace
doi:10.2514/2.4961 and Electronic Systems, Vol. 44, No. 1, 2008, pp. 4156.
[18] Li, R. R., and Jilkov, V. P., Survey of Maneuvering Target Tracking, doi:10.1109/TAES.2008.4516988
Part V. Multiple Model Methods, IEEE Transactions on Aerospace and [38] Nicholas, R. A., and Reichert, R. T., Gain Scheduling for H-Infinity
Electronic Systems, Vol. 41, No. 4, 2005, pp. 12551321. Controllers: A Flight Control Example, IEEE Transaction on Control
doi:10.1109/TAES.2005.1561886 Systems Technology, Vol. 1, No. 2, 1993, pp. 6979.
[19] Bowran, C., Artificial Neural Network Approaches to Target Recog- doi:10.1109/87.238400
nition, Digital Avionic System Conference, AIAA, Washington, D.C., [39] Gutman, O., and Palmor, Z. J., Proportional Navigation Against
Oct. 1988, pp. 847857. Multiple Targets, Journal of Guidance, Control, and Dynamics,
[20] Duh, F. B., and Lin, C. T., Tracking a Maneuvering Target Using Vol. 34, No. 6, 2011, pp. 17281733.
Neural Fuzzy Network, IEEE Transactions on Systems, Man, and doi:10.2514/1.53373
Cybernetics, Part B, Cybernetics, Vol. 34, No. 1, 2003, pp. 1633. [40] Palumbo, N. F., Blauwkamp, R. A., and Lloyd, J. M., Basic Principles
doi:10.1109/TSMCB.2003.810953 of Homing Guidance, Johns Hopkins APL Technical Digest, Vol. 29,
[21] Oshman, Y., and Arad, D., Enhanced Air-to-Air Missile Tracking No. 1, 2010, pp. 2541.
Using Target Orientation Observations, Journal of Guidance, Control, [41] Shima, T., Idan, M., and Golan, O. M., Sliding-Mode Control for
and Dynamics, Vol. 27, No. 4, 2004, pp. 595606. integrated Missile Autopilot Guidance, Journal of Guidance, Control,
doi:10.2514/1.11155 and Dynamics, Vol. 29, No. 2, 2006, pp. 250260.
[22] Atir, R., Hexner, G., and Weiss, H., Target Maneuver Adaptive doi:10.2514/1.14951
Guidance Law for a Bounded Acceleration Missile, Journal of

You might also like