Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

JOURNAL OF GUIDANCE, CONTROL, AND DYNAMICS

Vol. 29, No. 2, March–April 2006

Sliding-Mode Control for Integrated Missile Autopilot Guidance


Tal Shima∗
RAFAEL Armament Development Authority, Ltd., 31021 Haifa, Israel
Moshe Idan†
Technion—Israel Institute of Technology, 32000 Haifa, Israel
and
Oded M. Golan‡
RAFAEL Armament Development Authority, Ltd., 31021 Haifa, Israel

A sliding-mode controller is derived for an integrated missile autopilot and guidance loop. Motivated by a
differential game formulation of the guidance problem, a single sliding surface, defined using the zero-effort miss
distance, is used. The performance of the integrated controller is compared with that of two different two-loop
Downloaded by SIMON FRASER UNIVERSITY on October 9, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.14951

designs. The latter use a sliding-mode controller for the inner autopilot loop and different guidance laws in the
outer loop: one uses a standard differential game guidance law, and the other employs guidance logic based on
the sliding-mode approach. To evaluate the performance of the various guidance and control solutions, a two-
dimensional nonlinear simulation of the missile lateral dynamics and relative kinematics is used, while assuming
first-order dynamics for the target evasive maneuvers. The benefits of the integrated design are studied in several
endgame interception engagements. Its superiority is demonstrated especially in severe scenarios where spectral
separation between guidance and flight control, implicitly assumed in any two-loop design, is less justified. The
results validate the design approach of using the zero-effort miss distance to define the sliding surface.

Nomenclature γ = flight-path angle


A, B, C, G = state-space model matrices  = modeling error
a = acceleration T = maneuver period
D = drag force φ = maneuver phase
f (·) (·) = missile nonlinear aerodynamics function δ = canard deflection angle
I = moment of inertia θ = pitch attitude angle
K = controller gain λ = angle between the temporary and initial LOS
L = lift force μ = uncertainty control gain
L (·) = lift force derivative σ = sliding variable
 = Lyapunov function τ = time constant
M = pitch moment = transition matrix
M(·) = pitch-moment derivative ψ = time-varying function
m = mass
q = pitch rate Subscripts
r = range a = acceleration
T = engine thrust bf = body coordinates parallel to the inertial frame I
t = time br = rotating body-fixed coordinates
tf = final time G = guidance
tgo = time to go GC = integrated guidance control
V = speed I = inertial coordinate frame
x = state vector M = missile
Z = zero-effort miss MC = maneuvering command
z = target-missile relative displacement normal to the N or λ = normal to the LOS
initial line of sight (LOS) q = pitching moment
α = angle of attack r = radial, along the LOS
s = servo actuator
Presented as Paper 2004-4884 at the Guidance Navigation and Control T = target
Conference, Providence, RI, 16–19 August 2004; received 6 December 0 = initial values
2004; revision received 4 May 2005; accepted for publication 6 May 2005.
Copyright  c 2005 by the authors. Published by the American Institute of
Superscripts
Aeronautics and Astronautics, Inc., with permission. Copies of this paper
may be made for personal or internal use, on condition that the copier pay B = body contribution to the aerodynamics force
the $10.00 per-copy fee to the Copyright Clearance Center, Inc., 222 Rose- and moment
wood Drive, Danvers, MA 01923; include the code 0731-5090/06 $10.00 in c = command
correspondence with the CCC.
∗ System Engineer; currently National Research Council Visiting Scien- eq = equivalent control
tist, Air Vehicles Directorate, Air Force Research Laboratories, Room 304, max = maximum
Building 146, Wright–Patterson AFB, OH 45433; shima tal@yahoo.com. (¯) = approximation/model
Senior Member AIAA.
† Associate Professor, Faculty of Aerospace Engineering; moshe.idan@
I. Introduction
technion.ac.il. Associate Fellow AIAA.
‡ Chief Systems Engineer, P.O. Box 2250, Department 35; odedgol@
rafael.co.il. Senior Member AIAA. F LIGHT control and guidance loops of interceptor missiles are
usually assumed to be spectrally separated. Thus, a hierarchical
250
SHIMA, IDAN, AND GOLAN 251

design approach is commonly used, in which an inner-loop autopi- task. In this section, the SMC methodology and its previous use
lot is constructed to follow acceleration commands issued by the in missile guidance and control are first briefly reviewed. Then we
outer-loop guidance algorithm. The guidance law is designed based address the central aspect of this work: the proper definition of the
on a low-order approximation of the autopilot dynamics. Such a sliding surface that dictates the SMC design.
practice allows simple, lower-order, analysis of each of the design
cycles and enables conventional gain scheduling.1 However, close A. SMC—Brief Review
to interception spectral separation might not be valid because of The SMC methodology is described in many papers and text-
rapid changes in the relative geometry. This can cause instability books including Refs. 7 and 9 and will only briefly be reviewed
and, consequently, unacceptable miss distances. in this section. SMC is a robust control design approach enabling
A possible development in future interceptors is the requirement to maintain stability and performance in the presence of modeling
for very small miss distances in order to reduce warheads’ size (and errors. Simplified controllers are obtained using SMC by convert-
consequently their lethal radii) without affecting the interception ing a tracking problem of an nth order dynamical system into a
effectiveness. The integrated flight control and guidance law de- first-order stabilization problem. This approach leads to satisfactory
sign can enhance the endgame performance of the interceptor by performance in the presence of bounded but otherwise arbitrary pa-
accounting for the coupling between the control and guidance dy- rameter inaccuracies and model uncertainties. This is achieved by
namics. The integrated design that provides synergism between the large magnitude, often high-frequency, control activity that can be
two control components can also postpone the endgame instability. at odds with excitation of neglected high-frequency dynamics. To
In such a design, the entire guidance and control loop is stated as overcome this behavior, the design procedure involves tradeoffs be-
a solution to a finite horizon control problem, instead of the com- tween modeling and performance that are quantified in a simple and
Downloaded by SIMON FRASER UNIVERSITY on October 9, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.14951

mon approach treating the inner autopilot loop as an unrealistically intuitive fashion.
infinite horizon one. The integrated design also allows for a more The SMC design is performed around a sliding surface or man-
effective use of the information on the missile states in the guidance ifold, commonly denoted by σ = 0, where the sliding variable σ is
problem formulation, as opposed to using only the missile acceler- a function of the system tracking error and possibly its derivatives.
ation data in separated guidance loop designs. The number of derivatives used to define the sliding surface is dic-
The potential for improved performance of an integrated tated by the relative degree of the system. The sliding surface is
autopilot-guidance design has motivated a considerable research defined so that if the tracking error is confined to that surface this
effort in this area during the last decade. In Ref. 2, a game theoretic error will decay to zero with desired dynamics. The problem is to
approach was used for the design of an integrated autopilot-guidance drive the scalar quantity defining the sliding surface to zero and
linear controller, which minimizes the final miss distance and con- maintain it there, ultimately achieving exact tracking. The control
trol energy under worst-case target maneuvers and measurements action required for this task, obtained by imposing σ̇ = 0 once on
uncertainty. The feedback linearization method3 in conjunction with the surface, is denoted as the equivalent control. When the system
the linear-quadratic-regulator approach were used in Ref. 4 for the response is confined to the sliding surface, it is said that the system
design of an integrated controller. A similar approach was used in is in a sliding mode.
Ref. 5 in a finite horizon problem setting. In the latter, numerical Erroneous equivalent control values can be obtained as a result
nonlinear solution techniques have been employed requiring an on- of system uncertainties and disturbances. To accommodate the re-
line solution of a two-point boundary-value problem. In Ref. 6, a sulting departures from the sliding surface, the equivalent controller
state-dependent Riccati differential equation approach was used for is augmented by a second component, sometimes referred to as the
designing the integrated controller. Using a six-degrees-of-freedom uncertainty controller, the goal of which is to drive the system to
simulation, it was shown that the integrated controller provides im- the sliding surface in finite time, while ensuring closed-loop sta-
proved miss distance statistics compared to the conventional two- bility. The control laws that satisfy the stability requirement, while
loop design practice. assuming only the knowledge of the uncertainty bounds, are usu-
The sliding-mode-control (SMC) methodology7 is an intuitive ally discontinuous across the sliding surface. This discontinuity can
and simple robust control technique, addressing highly nonlinear cause high-frequency control chattering. Although in some applica-
systems with large modeling errors. SMC was successfully used in tions chattering might be acceptable, it is often required to limit the
the design of controllers for the autopilot and guidance loops, as control activity even by compromising closed-loop performance.
will be outlined in the next section. The main contribution of the This is achieved by smoothing the discontinuous components of the
current work is in the derivation of an integrated autopilot-guidance uncertainty controller or by introducing a boundary layer around the
controller using the SMC approach, utilizing the zero-effort miss sliding surface.
distance as its single sliding surface. The SMC design methodology entails three major steps: 1) selec-
A difficulty arises when applying SMC to nonminimum phase tion of a sliding surface σ = 0 (or surfaces in a multi-input/multi-
(NMP) systems because of the implicit plant inversion of the output problem) to ensure stable desired dynamic characteristics of
method. Different ways were suggested to treat NMP plants, for the system once in the sliding mode; 2) computation of the equiva-
example, in Ref. 8, by introducing dynamic sliding manifolds. In lent control to impose σ̇ = 0 once on the sliding surface, while using
this paper we examine a canard control configuration that has min- an approximate model of the system dynamics; and 3) choosing an
imum phase acceleration dynamics. This type of airframe is often uncertainty controller to ensure stability and finite time convergence
used in short-range air-to-air missiles because of its high agility to the surface when σ = 0.
compared to a tail control configuration.
The remainder of this paper is organized as follows: In the next B. SMC for Missile Autopilot and Guidance
section, the methodology of designing autopilot and guidance con- The SMC methodology can be used to construct missile autopilot
trollers using the SMC approach is discussed. Then, the nonlin- and guidance algorithms. In Ref. 10, a method was introduced for
ear engagement kinematics and missile dynamics equations are re- optimally selecting a time-varying sliding surface for a recursive
viewed along with the simplified models used for the autopilot and linear time-varying dynamics approximation of nonlinear system
guidance law designs. This is followed by the synthesis of an in- dynamics. The method was used to design an autopilot that opti-
tegrated autopilot-guidance control law, along with two different mally follows a given acceleration command. The sliding surface
conventional two-loop designs presented for comparison. A perfor- was composed of the angle of attack, pitch rate, and velocity errors.
mance analysis is then presented followed by concluding remarks. A high-order SMC for a robust missile autopilot design was pre-
The system zero dynamics are analyzed in the Appendix. sented in Ref. 11. The design was based on the relative degrees of
the plant and actuator dynamics, whereas the controller robustness
II. Methodology was aimed at addressing uncertainties in the nonlinear model for
In this paper the SMC methodology is used to design guidance these dynamics. A multi-input/multi-output problem was treated in
and control algorithms for the endgame phase of an interception Ref. 12 resulting in multiple sliding manifolds. A 180-deg heading
252 SHIMA, IDAN, AND GOLAN

reversal maneuver was sought while using different manifolds for


different stages of the maneuver. The goal was to cope with the vary-
ing missile dynamics resulting from the varying flight conditions.
A missile guidance law in the class of proportional navigation
(PN), derived using the SMC approach, was proposed in Ref. 13. The
sliding surface was selected to be proportional to the line-of-sight
(LOS) rate, and the target maneuvers were considered as bounded
uncertainties. Using numerical simulations, the superiority of the
proposed guidance law over the conventional PN was advocated. In
Ref. 14 an adaptive sliding-mode guidance law was derived. Using
analysis and simulations, robustness to disturbances and parameter
perturbations was shown.
A SMC autopilot-guidance controller was recently presented in
Ref. 15. It is composed of a two-loop design, using backstepping
and high-order SMC methods. In the outer SMC-like guidance loop,
a sliding surface that depends on the LOS rate was defined with Fig. 1 Planar engagement geometry.
the missile pitch rate used as a virtual control. The inner loop was
designed to robustly enforce the pitch-rate command of the outer
loop in the presence of uncertainties. Numerical simulation was used
Downloaded by SIMON FRASER UNIVERSITY on October 9, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.14951

to demonstrate the performance and robustness of the integrated


design in tracking an evasive maneuvering target, in the presence of
atmospheric disturbances and uncertainty in the plant and actuator
Fig. 2 Missile coordinate systems.
dynamics.

C. Zero-Effort Miss Sliding Surface


The selection of a sliding surface is a central part of the sliding-
mode controller design process. As just mentioned, the LOS rate is
commonly used to define the sliding surface because the underlying
physical principle to be sought is clear: attaining a collision triangle. A. Nonlinear Kinematics and Dynamics
If the speeds of both the interceptor and the target are constant, and if 1. Engagement Kinematics
they are on a collision course, then the LOS between them does not In Fig. 1 a schematic view of the planar endgame geometry is
rotate. To attain a collision course, the angular velocity command of shown, where X I − O I − Z I is a Cartesian inertial reference frame.
the interceptor missile can be chosen to be proportional to the LOS We denote the missile and target by the subscripts M and T , re-
rate, leading to the common PN guidance law. spectively. The speed, normal acceleration, and flight-path angles
Another natural candidate for defining the sliding surface is the are denoted by V , a, and γ , respectively; the range between the
well-known zero-effort miss distance16 (ZEM) commonly used in adversaries is r , and λ is the angle between the LOS and its initial
advanced optimal control and differential game formulations of the direction.
guidance problem. In a one-sided optimal control optimization prob- Neglecting the gravitational force, the engagement kinematics,
lem it has the physical meaning of being the miss distance, if from expressed in a polar coordinate system (r, λ) attached to the missile,
the current time onwards the interceptor does not apply controls and is
the target performs the expected maneuver. In a two-sided differ-
ential game problem, ZEM is the miss distance, if from the current ṙ = Vr (1a)
time onwards both players do not apply controls. This scalar vari-
able is obtained from the homogenous solution of the associated λ̇ = Vλ /r (1b)
engagement equations of motion. Generally, the ZEM depends on
the LOS rate and on the missile and target maneuvers. It can also where the closing speed Vr is
depend on additional terms stemming from varying speeds of both
Vr = −[VM cos(γ M − λ) + VT cos(γT + λ)] (2)
adversaries.17
Using the differential game-based ZEM to define the sliding sur-
and the speed perpendicular to the LOS is
face has several advantages. First, if the system response is main-
tained on the sliding surface, it provides zero miss distance. Once Vλ = −VM sin(γ M − λ) + VT sin(γT + λ) (3)
on the surface, no control action is needed to ensure interception in
the nominal case with perfect modeling, as long as the target does We approximate the time to go tgo that will be used in further
not issue maneuvering commands. In addition, it reduces the n- derivations by
dimensional guidance problem to a scalar one, with a dynamic equa-
tion that in the nominal case depends only on the system inputs.18 tgo = −r/Vr (4)
Therefore, ZEM is chosen to define the sliding surface for the sep-
arated guidance and integrated autopilot-guidance designs. 2. Target and Missile Dynamics
During the endgame, the target is assumed to move at a constant
III. Model Derivation speed. In addition we assume first-order lateral maneuver dynamics.
A skid-to-turn cruciform canard controlled missile, which is roll This is expressed by
stabilized, is considered. The motion of such a missile can be sep-  
arated into two perpendicular channels. During the end-game the ȧT = aTc − aT τT (5a)
missile is assumed to fly close to a collision course and the guid-
ance and control problem can be treated as planar in each of these γ̇T = aT /VT (5b)
channels. We first present the full nonlinear kinematics and dynam-
ics equations of the interception problem, which will serve for the where τT is the time constant of the target dynamics and aTc is the
synthesis and analysis of the nonlinear sliding mode guidance and maneuver command.
control designs presented in the sequel. Then, linearized equations The missile planar dynamics are expressed using the coordi-
will be derived, which will serve as the basis for the selection of the nate systems presented in Fig. 2. Here, X br − M − Z br is a rotat-
various sliding surfaces. ing body-fixed coordinate frame with the X br axis aligned with the
SHIMA, IDAN, AND GOLAN 253

missile’s longitudinal axis; X bf − M − Z bf is parallel to the inertial


frame X I − O I − Z I , with its origin attached to the missile center
of gravity.
Let α and θ denote the missile’s angle of attack and its pitch
attitude angle, respectively. Then, the following relation holds:
θ = α + γM (6)

Using the preceding definitions, the planar missile dynamics can be


expressed as
V̇M = [T cos α − D(α, δ)]/m (7a)

α̇ = q − [T sin α + L(α, δ)]/(mVM ) (7b)

q̇ = M(α, q, δ)/I (7c)

θ̇ = q (7d)

δ̇ = (δ − δ)/τs
c
(7e) Fig. 3 Linearized endgame kinematics.
Downloaded by SIMON FRASER UNIVERSITY on October 9, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.14951

where q is the missile pitch rate; m and I are its mass and mo-
ment of inertia, respectively; δ is the deflection angle of a canard where the subscript 0 denotes the initial value around which lin-
controlled by a servo actuator represented by a first-order dynamics earization has been performed.
model with a time constant τs ; M is the pitch moment acting on the For nonintegrated guidance law designs, we assume that the
missile; T is the engine thrust that is aligned with X br axis; and L closed-loop maneuver dynamics of the missile can be approximated
and D are, respectively, the lift and drag forces. The aerodynamic by an equivalent first-order system with a time constant τ M . There-
forces and moments are nonlinear, partly unknown functions of the fore, the equations of relative motion normal to the initial LOS are
related variables, in particular α, δ, and q. It is assumed that during expressed by
the endgame the missile has no thrust, and its speed variation is
negligible. Hence, the model in Eqs. (7) is reduced to a fourth-order ẋG = A G xG + BG aMN
c
+ G G aTN
c
(13)
model given by
where
α̇ = q − L(α, δ)/(mVM ) (8a)
⎡ ⎤
  0 1 0
q̇ = M(α, q, δ)/I (8b) A G11 A G12
AG = , A G11 ⎣
= 0 0 1 ⎦
[0]1 × 3 −1/τ M
θ̇ = q (8c) 0 0 −1/τT

δ̇ = (δ c − δ)/τs (8d)
⎡ ⎤
0
A G12 = ⎣−1⎦ (14)
B. Linearized Kinematics and Dynamics
Because of the complexity of the preceding nonlinear models, the 0
definition of the sliding surfaces in the various SMC guidance and ⎡ ⎤ ⎡ ⎤
control designs will be based on simplified kinematics and dynam- 0 0
ics models. We will assume that during the endgame the missile and ⎢ 0 ⎥ ⎢ 0 ⎥
target deviations from the collision triangle are small. Hence, lin- BG = ⎢
⎣ 0 ⎦,
⎥ GG = ⎢ ⎥
⎣1/τT ⎦ (15)
earization of the endgame kinematics can be performed around the
1/τ M 0
initial LOS.16 For the control problem we will use the approximate
short-period linearized equations of motion.1 c c
and aTN and aMN are, respectively, the target and missile acceleration
1. Engagement Kinematics
commands normal to the initial LOS. In Eq. (14) and hereafter, [0]
denotes a matrix of zeros with appropriate dimensions.
The linearized endgame kinematics used for the decoupled guid-
ance law design is depicted in Fig. 3, where the X axis is aligned
with the initial LOS. With the assumption that during the endgame 2. Missile Dynamics
the speeds of the missile and target are constant, the closing speed Vr The linearized actuator and short-period missile dynamics, as-
is constant, and the interception time, given by t f = −r0 /Vr , can be suming constant speed, are expressed using the state-space vector
assumed fixed. Therefore, the time to go of Eq. (4) can be computed
as x M = [α q δ]T (16)

tgo = t f − t (9) Following Eqs. (8), these dynamics are given by


The state vector of the linearized guidance problem is defined by
ẋ M = A M x M + B M δ c (17)
xG = [z ż aTN aMN ] T
(10)
where
where z is the relative displacement between the target and the ⎡ ⎤ ⎡ ⎤
missile normal to the initial LOS direction. The target and missile −L α /VM 1 −L δ /VM 0
accelerations normal to the LOS are denoted by aTN and aMN , re- AM = ⎣ Mα Mq Mδ ⎦ , B M = ⎣ 0 ⎦ (18)
spectively, and, under the linearization assumption, satisfy
0 0 −1/τs 1/τs
aMN ≈ a M cos(γ M0 − λ0 ) (11)
and L (·) and M(·) denote the dimensional stability and control deriva-
aTN ≈ aT cos(γT 0 + λ0 ) (12) tives of the short-period model.
254 SHIMA, IDAN, AND GOLAN

The missile acceleration normal to the initial LOS, which will Differentiating Eq. (29) with respect to time yields
serve as the control variable in the outer guidance loop of the two-
loop guidance and control designs, is given by z + żtgo = −Vr tgo
2
λ̇ (30)

aMN = C M x M (19) Using this expression, the guidance ZEM of Eq. (26) can be ex-
pressed as
where
Z G = −Vr tgo
2
λ̇ + aTN τT2 ψ(tgo /τT ) − aMN τ M2 ψ(tgo /τ M ) (31)
C M = [L α 0 L δ ] cos(γ M0 − λ0 ) (20)
Note that although λ̇ and aMN are expected to be available through
3. Integrated Dynamics
measurements, the target acceleration aTN must be estimated.
The state vector of the integrated autopilot-guidance problem is
defined by
2. ZEM for Integrated Autopilot Guidance
xGC = [z ż aTN α q δ]T (21) The preceding procedure can now be used to determine the ZEM
for the integrated autopilot-guidance dynamics of Eqs. (22–24),
The equations of motion for this case are leading to
ẋGC = AGC xGC + BGC δ c + G GC aTN
c
(22) 
Z GC = z + żtgo + aTN τT2 ψ(tgo /τT ) + ψα (tgo )α + ψq (tgo )q + ψδ (tgo )δ
Downloaded by SIMON FRASER UNIVERSITY on October 9, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.14951

where (32)
⎡ ⎤
  [0]1 × 3 where ψα (tgo ), ψq (tgo ), and ψδ (tgo ) are complicated functions of the
A G11 A12
AGC = , A12 = −C M ⎦
⎣ (23) system parameters and tgo and hence of the kinematics variables. Al-
[0]3 × 3 AM ternatively, the last three terms of Eq. (32) can be obtained from the
[0]1 × 3 numerically computed transition matrix GC (tgo ) associated with
AGC of Eq. (23). Using this approach and applying the relationship
BGC = [[0]1 × 5 1/τs ]T , G GC = [0 0 1/τT 0 0 0]T given in Eq. (30), the ZEM is given by
(24) Z GC = −Vr tgo
2
λ̇ + aTN τT2 ψ(tgo /τT ) + CGC GC (tgo )x̄GC (33)
with A G11 , A M , and C M already defined in Eqs. (14), (18), and (20), where
respectively.
GC (tgo ) = exp(AGC tgo ) (34)
C. Zero-Effort Miss
As discussed in Sec. II, ZEM is used in the various SMC solu- CGC = [1 0 0 0 0 0] (35)
tions considered in this paper. It depends on the interception time (or
more precisely on the time to go) and is obtained from the homoge- x̄GC = [0 0 0 α q δ] T
(36)
nous solution of the equations of motion. Therefore, ZEM depends
on the analyzed problem and the related model. For complex non- IV. Autopilot and Guidance Synthesis
linear interception models ZEM cannot be determined analytically. To demonstrate the advantages of the proposed integrated guid-
Therefore, in the current work approximate ZEMs are derived for ance and control solution, two additional designs are presented,
linearized models. Nonetheless, the dynamic characteristics of the in which the guidance and flight control problems are addressed
ZEMs will be determined using the nonlinear models. This approx- separately. All three designs use the SMC methodology in their
imation approach will be validated in Sec. V through simulation. various components and use the ZEM in the guidance loop. The
first solution, representing the baseline design, couples a differen-
1. ZEM for Guidance tial game law (DGL) with an autopilot designed using the sliding
For the separated guidance problem, the governing equations are mode approach. This design will be referred to as DGL-SMC in the
(13–15). Following Ref. 18, let G (t f , t) be the transition matrix sequel. Then, in the second solution the former SMC autopilot is
associated with these dynamics. For the investigated time-invariant commanded by a guidance law constructed using the SMC method-
problem it can be computed as ology. This design will be termed SMG-SMC. Finally, the proposed
integrated guidance and control solution, SMGC, is presented.
G (t f , t) = G (tgo ) = exp(A G tgo ) (25) For the synthesis performed in this section and the analysis per-
formed next, we assume that perfect information on the states of the
The sparse structure of A G allows for an analytical solution of problem is available. Thus, if a state is not measurable, such as the
G (tgo ),
which is then used to determine the guidance ZEM as target acceleration aT , we assume it can be estimated.

Z G = CG G (tgo )xG = z + żtgo + aTN τT2 ψ(tgo /τT )
A. DGL-SMC Baseline Design
− aMN τ M2 ψ(tgo /τ M ) (26) In this baseline design, the DGL is used in the guidance loop. It
was derived for the approximate linearized model of the endgame
where kinematics and first-order target and missile dynamics given by
Eqs. (13–15), while assuming bounded target and missile accel-
C G = [1 0 0 0] (27) eration commands.19 This guidance law is given by
and c
aMN = aMN
max
sgn(Z G ) (37)

ψ(ζ ) = exp(−ζ ) + ζ − 1 (28) c
where Z G is the ZEM given by Eq. (31), and a M , acting as a com-
Alternatively, the first two terms in Eq. (26) can be expressed as a mand to the inner-loop autopilot, is given by
function of the kinematics variables Vr and λ. Specifically, using 
the assumption of small deviations from a collision triangle, the
c
aM = aMN
c
cos(γ M − λ) (38)
displacement z normal to the initial LOS can be approximated by
We assume that during the endgame |γ M − λ| < π/2. Note that aMN
max

z ≈ (λ − λ0 )r (29) max
relates to a M by the same relation as in Eq. (38).
SHIMA, IDAN, AND GOLAN 255

The missile autopilot is derived using the SMC methodology. We Substituting Eqs. (50) into (49) leads to
assume that the lift and the aerodynamic pitch moment in Eqs. (8)
are comprised of the body and fin contributions. This is modeled ā˙ M = K α [q − (ā M + a )/VM ] + K δ (δ c − δ) (51)
by
where
L/m = L αB f 1 (α) + L δ f 2 (α + δ) (39a)
K α = L̄ αB f¯1 (α) + L̄ δ f¯2 (α + δ) (52a)
M/I = MαB f 3 (α) + Mq q + Mδ f 4 (α + δ) (39b)
K δ = L̄ δ f¯2 (α + δ)/τs (52b)
where
The SMC controller is chosen as
L αB = L α − L δ (40a)
δ c = δ eq − μ sgn(σ )/K δ (53)
MαB = Mα − Mδ (40b)
where δ eq is the equivalent controller given by
and f i (·), i = 1, 2, 3, 4 are bounded functions expressing the non-
linear aerodynamic characteristics of the missile. The autopilot is δ eq = δ − (K α /K δ )(q − ā M /VM ) (54)
designed to control the normal acceleration
and the second term on the right-hand side of Eq. (53) is the uncer-
a M = L/m (41) tainty controller. With this controller, the derivative of the Lyapunov
Downloaded by SIMON FRASER UNIVERSITY on October 9, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.14951

function candidate becomes


with L/m defined in Eq. (39a). 
The SMC autopilot is designed using an approximation of the ˙ = −σ μ sgn(σ ) + K α a /VM + ȧ c
 (55)
M
nonlinear model of Eqs. (8), (39), and (40), that is,
 Assuming that the derivative of the acceleration command is
α̇ = q − L̄ αB f¯1 (α) + L̄ δ f¯2 (α + δ) VM (42a) bounded by
 c
ȧ M  ≤ 
¯ MC
q̇ = M̄αB f¯3 (α) + M̄q q + M̄δ f¯4 (α + δ) (42b) (56)

and using the bound in Eq. (45), the derivative in Eq. (55) can be
δ̇ = (δ c − δ)/τs (42c)
bounded as
and ˙ ≤ −|σ |[μ − |K α /VM |
 ¯a −
¯ MC ] (57)
ā M = L̄ αB f¯1 (α) + L̄ δ f¯2 (α + δ) (43)
Choosing
where L̄ (·) , M̄(·) , and f¯i (·), i = 1, 2, 3, 4 are approximations of their ¯a +
¯ MC
μ > |K α /VM | (58)
respective quantities and functions. It is assumed that the difference
between a M and ā M ensures that  ˙ is negative definite and that the sliding surface σ = 0
a M − ā M = a (44) of Eq. (46) is reached in a finite time. This, in turn, implies that
perfect tracking of the commanded acceleration can be attained in
is bounded a finite time.
To comply with the condition in Eq. (56), the discontinuous guid-
¯a
|a | ≤  (45) ance command of Eq. (37) is smoothed by a high bandwidth low-
pass filter. In addition, to avoid chattering as a result of the discon-
A canard-controlled missile with normal acceleration measure- tinuous command in Eq. (53), a boundary-layer approximation of
ment is minimum phase. Moreover, the normal acceleration output the sign function can be used, thus compromising exact tracking by
has a relative degree of one with respect to the actuator command uniform ultimate boundedness of the tracking error.9
δ c . Therefore, the SMC sliding surface, σ = 0, is defined using the
variable B. SMG-SMC Design
σ = ā M − c
aM (46) The decoupled guidance and control design just presented is mod-
ified by redesigning its guidance law using the SMC approach. This
The SMC autopilot is designed using the following Lyapunov is mainly aimed at testing the validity of using the ZEM in defining
function candidate: the guidance sliding surface. The guidance law design is based on
the ZEM of Eq. (31). Although the ZEM was computed using a
 = 12 σ 2 (47) linearized model, the consequent guidance logic, which involves a
time derivative of the ZEM, will be determined using the nonlinear
The time derivative of this Lyapunov function candidate is model.
  As will be evident from the subsequent derivation, the ZEM,
˙ = σ σ̇ = σ ā˙ M − ȧ c
 (48) which we want to regulate to zero, has a relative degree of one with
M
c
respect to the guidance control signal aMN . Therefore, the variable
where chosen to define the guidance sliding surface is
 
ā˙ M = L̄ αB f¯1 (α) α̇ + L̄ δ f¯2 (α + δ) (α̇ + δ̇) (49) σG = Z G (59)

and f¯i (·), i = 1, 2 denote the partial derivatives of the functions The stability of this SMC guidance law is ensured through the
f¯i (·), i = 1, 2 with respect to their respective arguments. Using the analysis of the following candidate Lyapunov function:
modeling error relation of Eq. (44) in the nonlinear model of Eqs. (8),
the state derivatives in Eq. (49) can be expressed as G = 12 σG2 (60)

α̇ = q − a M /VM = q − (ā M + a )/VM (50a) Its time derivative is given by

δ̇ = (δ c − δ)/τs (50b) ˙ G = σG Ż G
 (61)
256 SHIMA, IDAN, AND GOLAN

where Using Eqs. (70) and (71), the time derivative of the guidance
Lyapunov function candidate of Eq. (61) can be bounded as
Ż G = −(V̇r tgo λ̇ + 2Vr t˙go λ̇ + Vr tgo λ̈)tgo + [ȧTN ψ(tgo /τT )
˙ G ≤ −|σG |(μG − 
 ¯ aTNc − 
¯ aTNτ − 
¯ aMNτ ) (72)
+ aTN ψ  (tgo /τT )t˙go ]τT2 − [ȧMN ψ(tgo /τ M )
Setting
+ aMN ψ  (tgo /τ M )t˙go ]τ M2 (62)
¯ aTNc + 
μG >  ¯ aTNτ + 
¯ aMNτ (73)
In the latter,
  ensures that the guidance sliding surface is attained in finite time.
tgo ∂ψ(tgo /τT ) tgo /τT − ψ(tgo /τT )
ψ = = (63) The system response while in sliding mode, that is, its zero dynam-
τT ∂tgo τT ics, is investigated in the Appendix. In a realistic noisy interception
environment, a boundary layer around the ZEM sliding surface can
with a similar result for ψ  (tgo /τ M ). Based on the constant target and be employed to provide smooth acceleration commands.
missile speed assumptions and using the engagement kinematics of
Eqs. (1–4), the derivatives V̇r , λ̈ and t˙go are given by C. SMGC Design
 For the integrated guidance and control design, the full nonlinear
V̇r = Vλ r + a M sin(γ M − λ) + aT sin(γT + λ)
2
(64a) kinematics and dynamics model is used. The model input is the
actuator command δ c , while the target acceleration command aTN c
V̇λ = −Vλ Vr /r − a M cos(γ M − λ) + aT cos(γT + λ) (64b)
Downloaded by SIMON FRASER UNIVERSITY on October 9, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.14951

is treated as a disturbance. Similar to the SMG-SMC design, the


ZEM, this time of Eq. (33), will define the sliding surface. Here
λ̈ = V̇λ /r − Vλ Vr /r 2 (64c) again, the relative degree of this ZEM to the input is one, leading to
 the sliding variable
t˙go = −1 + V̇r r Vr2 (64d)
σGC = Z GC (74)
The true target and missile dynamics are assumed to be related
to the approximate first-order linear models by The candidate Lyapunov function and its time derivative are
 
ȧTN = aTN
c
− aTN τT + aTN (65a) GC = 12 σGC
2
(75)
  and
ȧMN = aMN
c
− aMN τ M + aMN (65b)
˙ GC = σGC Ż GC
 (76)
where
Recognizing that the first two terms of Z G in Eq. (31) and Z GC in
¯ aTN
|aTN | ≤  (66a) Eq. (33) are identical, the results of their differentiation obtained in
¯ aMN the SMG case just presented can be used here. Therefore, based on
|aMN | ≤  (66b) Eq. (33) and the results in Eq. (67),

are the bounded target and missile dynamics errors. Substituting Ż GC = {Vλ + aTN τT [1 − exp(−tgo /τT )]}V̇r r Vr2
Eqs. (63–65) into Eq. (62), after some simplifications, the ZEM
time derivative is manipulated to yield  
+ τT aTN
c
+ τT aTN ψ(tgo /τT ) − tgo aMN
Ż G = {Vλ + aTN τT [1 − exp(−tgo /τT )] − aMN τ M
   + CGC [  ˙
GC (tgo )x̄GC tgo + ˙
GC (tgo )x̄GC ] (77)
×[1 − exp(−tgo /τ M )]}V̇r r Vr2 + τT aTN
c
+ τT aTN ψ(tgo /τT )
where t˙go is given in Eq. (64d) and
 
− τ M aMN
c
+ τ M aMN ψ(tgo /τ M ) (67)  ∂ GC (tgo )
GC (tgo ) = = AGC GC (tgo ) (78)
Note that the first derivative of Z G yields the maneuver command ∂tgo
c
aMN ,asserting that the relative degree of the ZEM is one. Choosing
c
the control aMN as The derivatives α̇ and δ̇ in the vector x̄˙ GC were derived in Eqs. (50).
Using the relations in Eqs. (39b) and (42b), q̇ is expressed as
eq μG sgn(σG )
= aMN + q̇ = M̄αB f¯3 (α) + M̄q q + M̄δ f¯4 (α + δ) + q
c
aMN (68) (79)
τ M ψ(tgo /τ M )
eq Here q denotes a bounded error in the pitching-moment equation
where the equivalent control term aMN is set to satisfying
eq
 
aMN = V̇r r Vr2 τ M ψ(tgo /τ M ) {Vλ + aTN τT [1 − exp(−tgo /τT )] |q | ≤  ¯q (80)
− aMN τ M [1 − exp(−tgo /τ M )]} (69) Equations (50), (64d), (78), and (79) are used next to manipulate
Eq. (77) into
simplifies Eq. (67) to 
  Ż GC = {Vλ + aTN τT [1 − exp(−tgo /τT )] + CGC GC (tgo )ȳGC } V̇r r Vr2
Ż G = −μG sgn(σG ) + τT aTN
c
+ τT aTN ψ(tgo /τT )
(1,6)
 
+ GC (tgo )δ /τs
c
+ τT aTN
c
+ τT aTN ψ(tgo /τT ) + GC
− τ M2 ψ(tgo /τ M )aMN (70)
(81)
The last three unknown terms in Eq. (70) are assumed to be
bounded as where ȳGC = AGC x̄GC , āMN = −ȳGC (2) (minus the second element of
  ȳGC ), and (1,6)
GC (tgo ) is the (1,6)th element of the numerically com-
τT ψ(tgo /τT )aTN
c  ¯ aTNc
≤ (71a) puted transition matrix GC (tgo ), that is, the complicated function
 2  ψδ (tgo ) used in Eq. (32). All of the bounded modeling errors that
τT ψ(tgo /τT )aTN  ≤ 
¯ aTNτ (71b) were introduced to derive Eq. (81) are grouped into the bound GC .
Thus,
 2 
τ M ψ(tgo /τ M )aMN  ≤ 
¯ aMNτ (71c) |GC | ≤ ¯ GC (82)
SHIMA, IDAN, AND GOLAN 257

Motivated by Eq. (81), the SMGC controller is defined as L̄ αB = 1190 m/s2 , L̄ δ = 80 m/s2 , M̄αB = −234 s−2 , M̄q = −5 s−1 ,
 (1,6) M̄δ = 160 s−2 , a M max
= 40 g, and τ M = 0.1 s. The target parame-
δ c = δ eq − μGC sgn(σGC )τ S GC (tgo ) (83)
ters are VT = 380 m/s, aTmax = 20 g, T = 1 s, φ ∈ [0, 1] s, and
where the equivalent control is given by τT ∈ [0.05, 0.2] s. The approximate inner-loop parameters τ M and
    aM max
were obtained by evaluating the inner-autopilot-loop perfor-
−tgo mance when operating without the guidance loops. The functions
δ eq
= − Vλ + aTN τT 1 − exp
τT f i ( ), i = 1, 2, 3, 4 of the aerodynamic model [see Eqs. (39)] were
 chosen to be the standard saturation functions
V̇r r τs ⎧
+ CGC GC (tgo )ȳGC (1,6)
(84) ⎨Um Um < u
Vr2 GC (tgo ) sat(u) = u −Um ≤ u ≤ + Um
With this definition of the control signal δ c , the derivative in Eq. (76) ⎩
−Um u < − Um (87)
can be bounded by
˙ GC ≤ −|σGC |(μGC − 
¯ aTNc − 
¯ aTNτ − 
¯ GC ) with Um = 30 deg for all i. For the SMGC design, the model func-
 (85) tions f¯i ( ), i = 1, 2, 3, 4 were chosen to be identical to those in
By setting Eq. (87). However, because the SMC autopilot requires derivatives
of those functions with respect to their argument [see Eqs. (49) and
¯ aTNc + 
μGC >  ¯ aTNτ + 
¯ GC (86) (52)], a smooth approximation of Eq. (87) was used in the DGL-
the sliding surface Z GC = 0 can be reached in a finite time. The zero SMC and SMG-SMC designs. Note that all three designs have been
Downloaded by SIMON FRASER UNIVERSITY on October 9, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.14951

dynamics of the system is investigated in the Appendix. implemented with a boundary layer around the respective sliding
Similar to the autopilot SMC, chattering caused by the disconti- surfaces. In the inner loop of the DGL-SMC and SMG-SMC de-
nuity of the controller in Eq. (83) can be reduced by replacing the signs, the width of the boundary layer around the sliding surface
sign function by a high-gain linear element with saturation, that is, of Eq. (46) was chosen as 0.5 m/s2 , whereas no boundary layer
a boundary layer. was needed in the outer loop of SMG-SMC. The SMGC design
used a boundary layer of 5 cm around its sliding surface from
V. Performance Analysis Eq. (74).
Performance of the three guidance and control algorithms just
B. Sample Run Performance
developed is investigated in this section via numerical simulations,
using the nonlinear kinematics and missile and target dynamics of The performance of the various guidance and control designs is
Eqs. (1), (5), and (8). First, we outline the test scenario and its pa- first evaluated and compared for a sample run. The target maneuver
rameters. Then, we present a qualitative comparison between the is characterized by τT = 0.05 s and φ = 0.1 s. The sample run
three different designs using a sample run. A quantitative compar- was performed without missile model uncertainties. The missile
ison of the homing performance, conducted using a Monte Carlo aerodynamic parameters were chosen to be identical to the model.
simulation study, is then discussed. In Figs. 5–7 the ZEM, missile acceleration, and canard deflection
are compared for the three different designs.
A. Scenario
The ZEM plotted in Fig. 5 was computed using Eq. (33) for the
integrated design and using Eq. (31) for the other two designs. Large
The generic interceptor model used in this study is based on the
initial heading errors cause large initial ZEMs, reduced significantly
missile control example introduced in Ref. 20. It is assumed that the
at the beginning of the scenario for the three designs. Contrary to
target performs a square-wave evasive maneuver with a period of
the two-loop designs, the integrated SMGC solution does not exhibit
T and a phase of φ relative to the beginning of the simulation.
an overshoot in the ZEM and nearly nullifies it after the transient
The initial range was chosen as 1000 m. The engagement geometry
stage. The overshoot and somewhat oscillatory response of the DGL-
and trajectories are plotted in Fig. 4 for an example where the missile
SMC and SMG-SMC designs, triggered by the abrupt changes in
initial velocity vector is aligned with the initial LOS. The periodic
the target maneuvers, are attributed to the approximate first-order
maneuver of the target and consequently that of the missile are
modeling of the inner loop and its consequent effect on the outer-
evident. In addition, after a transient caused by the initial heading
loop performance.
error, the LOS orientation remains almost constant.
The missile acceleration for the three designs is plotted in
In this simulation study the constant missile speed is assumed
Fig. 6, depicting the periodic missile response imposed by the tar-
to be VM = 380 m/s. The canard servo is characterized by a
get maneuver. The acceleration profiles of the two-loop designs
time constant of τs = 0.02 s. The missile model parameters are

Fig. 4 Sample run engagement trajectories. Fig. 5 ZEM comparison for a sample run.
258 SHIMA, IDAN, AND GOLAN

a)
Downloaded by SIMON FRASER UNIVERSITY on October 9, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.14951

b)
Fig. 6 Acceleration profile comparison for a sample run.
Fig. 8 Homing performance comparison: a) expected miss and b) its
standard deviation.

Fig. 7 Canard deflection comparison for a sample run.

are similar throughout the maneuver, whereas that attained by the


SMGC solution is similar in magnitude and shape, but is slightly Fig. 9 SMGC sensitivity to modeling errors.
smoother.
The time history of the canard deflection is shown in Fig. 7. Dur- The sensitivity of the integrated design to modeling errors is de-
ing most of the maneuver, very oscillatory and for times saturated picted in Fig. 9. The lower bound on the homing performance is
canard deflections are observed with the two-loop controllers. Con- given by the results for the nominal case with no modeling errors. It
trary to this behavior, the SMGC controller led to much smoother is apparent that increasing the uncertainty interval has a nonlinear,
and smaller deflections, avoiding saturation during most of the ma- increasing effect on the homing performance.
neuver. This demonstrates one of the advantages of the integrated The performance of the SMGC controller was evaluated also for
design. varying initial conditions, such as changes in the initial heading
angles of the adversaries. It was found that the homing performance
C. Homing Performance was only slightly affected by those changes, resulting in similar
To evaluate and compare the homing performance of the dif- results to those presented in Fig. 8.
ferent algorithms, a Monte Carlo simulation study consisting of
100 sample runs for each test point was carried out. In these simula- VI. Conclusions
tions, for each test case the random variables were chosen to be the This paper demonstrates that by using an integrated design for
target maneuver phase φ and the missile aerodynamic parameters an interceptor missile autopilot and guidance loops, improved per-
(L αB , L δ , MαB , Mq , and Mδ ). These variables were assumed to be formance can be obtained compared to the conventional two-loop
distributed uniformly. design. The advantage of the integrated design is primarily signif-
In Fig. 8 the statistics (mean and standard deviation) of the miss icant in interception scenarios of highly maneuverable targets. In
distance are plotted for different dynamic capabilities of the tar- such cases, the inherent instability of the decoupled guidance and
get. The aerodynamics coefficients of the missile were varied ran- control loops near the terminal time is postponed by the integrated
domly by ±20% from the model values. It is apparent that the use design, resulting in reduced miss distances.
of the integrated guidance-control algorithm (SMGC) yields supe- The sliding-mode control approach has been used for the
rior homing performance, while the two-loop designs (DGL-SMC derivation of the integrated guidance-control algorithm. The zero-
and SMG-SMC) exhibit similar characteristics. This advantage is effort miss distance, obtained from a linearized differential game
amplified for highly maneuverable targets, characterized by small formulation of the interception problem, was used to define the
values of τT . For such targets the spectral separation between the sliding surface of the integrated design. This approach requires the
guidance and flight control is less justified. This demonstrates the selection of only one time-dependant controller gain, needed to en-
main advantage of the integrated design. sure convergence of the missile response to the sliding surface in
SHIMA, IDAN, AND GOLAN 259

the presence of model uncertainties and target maneuvers. Through The dynamics in Eqs. (A3) are not directly affected by the sys-
c
simulations it was shown that small miss distances can be achieved tem input aMN . Therefore, the system dynamics for the new states
even in stringent interception scenarios. are expressed in the normal form, where [r λ aT γT γ M ]T are its
The results of this study indicate that the integrated missile internal states.3 We now examine the boundedness of the internal
autopilot-guidance design, obtained by the sliding-mode approach states in our finite horizon problem, assuming we are on the sliding
with zero-effort miss used to define the sliding surface, has the poten- surface.
tial of improving interception accuracy in future endgame scenarios The target states are obtained by integrating the homogeneous
against highly maneuverable targets. The enhanced accuracy will Eq. (A3c) and then Eq. (A3d). This yields
enable the use of warheads with smaller kill radius and may even
eliminate their need. aT = exp(−t/τT )aT (0) (A5)

Appendix: Zero-Effort Miss Dynamics γT = [1 − exp(−t/τT )]aT (0)/τT + γT (0) (A6)


In this Appendix we analyze the closed-loop system behavior which are obviously bounded. Using the relations in Eqs. (2) and
when it is on the sliding surface, that is, its zero dynamics. In par- (3) together with the assumption that VM and VT are constant, we
ticular, we are interested to verify that while on the sliding surface, conclude that Vr and Vλ are bounded regardless of γT , γ M , and
that is, when the appropriate ZEM is identically zero, there are no λ. Consequently, integration of Eq. (A3a) over a finite time interval
unbounded internal states, and the control input is bounded. For leads to a bounded r . From Eq. (A3b) we conclude that λ̇ is bounded,
infinite horizon cases, this is ensured when the controlled system, except possibly when r → 0. It is easy to show that ψ(0) = 0 and
with the sliding variable defined as its output, is minimum phase.3 ψ  (ν) > 0 ∀ ν > 0. Thus we conclude that ψ(ν) > 0 ∀ ν > 0. Conse-
Downloaded by SIMON FRASER UNIVERSITY on October 9, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.14951

However, because the target interception scenario is a finite hori- quently, from Eq. (A4), aMN and a M are bounded, except for possibly
zon problem, boundedness of the system signals, while the ZEM is when r → 0.
identically zero, will be investigated without actually proving the Based on Eq. (A3e) it also implies that γ M is bounded for all t
system being minimum phase. except possibly when r → 0.
Now we examine the system response when r → 0. Using
Guidance ZEM Eq. (A4) we obtain that for r → 0 (or equivalently tgo → 0) and
In the two-loop autopilot and guidance design cases, the states while on the sliding surface
that define the guidance problem are x = [r λ aT γT γ M a M ]T . In
aMN (tgo → 0) ≈ −2Vr Vλ /r + aTN (A7)
this case Z G of Eq. (31) is used to define the guidance sliding surface.
The assumption that the system remains on the sliding surface once Substituting this result into Eq. (64b), we obtain
it is reached implies that there are no system uncertainties, modeling
errors, and unknown disturbances. If this was not the case, the system V̇λ (tgo → 0) ≈ Vr Vλ /r (A8)
response would have been moved away from the sliding surface as
a result of those uncertainties, contradicting the basic assumption Consequently, using Eq. (A8) in (64c) leads to
in the current analysis of remaining on the sliding surface.
The equations of motion of the system are given by Eqs. (1) λ̈(tgo → 0) ≈ 0 (A9)
and (5) and by first-order linear missile acceleration dynamics. To
Thus, because Eqs. (64c) and (A9) imply bounded λ̈ for all t, we
analyze the system dynamics on the sliding surface, we introduce
conclude that λ̇ and λ are bounded. Recalling Eqs. (64a) and (A3b),
the following state transformation T (·, t):
this implies also boundedness of aMN , a M , γ M , and V̇r .
y = T (x, t) = [r λ aT γT γM Z G ]T (A1) It remains to be shown that the commanded missile acceleration
required to remain on the sliding surface is bounded. This com-
In the new state variables, the last state Z G is the sliding variable, manded acceleration is obtained by setting Eq. (A2) to zero, result-
treated here as the system output. As shown earlier in Eq. (67), its ing in
 
relative degree is one. Because in this analysis the model uncertain- c
aMN = V̇r r Vr2 τ M ψ(tgo /τ M ) {Vλ + aTN τT [1 − exp(−tgo /τT )]
ties and disturbances are zero, Eq. (67) simplifies to
− aMN τ M [1 − exp(−tgo /τ M )]} (A10)
Ż G = {Vλ + aTN τT [1 − exp(−tgo /τT )] − aMN τ M
 All of the terms in the numerator of Eq. (A10) are bounded. We
×[1 − exp(−tgo /τ M )]}V̇r r Vr2 − aMN
c
τ M ψ(tgo /τ M ) (A2) have shown that ψ(ν) > 0 ∀ ν > 0, and because during the endgame
Vr < 0 we obtain that the commanded acceleration is also bounded,
c
where aMN is the system input. except possibly when r → 0. Using Eqs. (4) and (A3b), we obtain
The differential equations for the first five states in Eq. (A1), while that for r → 0, while on the sliding surface,
on the sliding surface and hence aTc = 0, are
c
aMN (tgo → 0) ≈ 2τ M V̇r λ̇ (A11)
ṙ = Vr (A3a)
Boundedness of λ̇ and V̇r imply also boundedness of the missile ac-
λ̇ = Vλ /r (A3b) celeration command. This completes the proof that all of the internal
states of the guidance problem and the acceleration command, when
ȧT = −aT /τT (A3c) solved using the SMG approach, are bounded when on the sliding
surface Z G (t) ≡ 0.
γ̇T = aT /VT (A3d)
Integrated ZEM
γ̇ M = a M /VM (A3e)
The states x = [r λ aT γT γ M α q δ]T define the integrated
autopilot-guidance problem. Here, Z GC of Eq. (33) is used to define
where Vr and Vλ are given by Eqs. (2) and (3), respectively.
the sliding surface. The equations of motion of the system are given
Based on Eq. (31), while on the sliding surface, the missile ac-
by Eqs. (1), (5), and (8).
celeration perpendicular to the LOS aMN is given by
We now introduce the following state transformation T (·, t):
 
aMN = −Vr tgo
2
λ̇ + aTN τT2 ψ(tgo /τT ) τ M2 ψ(tgo /τ M ) (A4) y = T (x, t) = [r λ aT γT γM α q Z GC ]T (A12)

and the total missile acceleration a M relates to aMN by the same In the new state variables, the last state Z GC is the sliding variable,
relation as in Eq. (38). treated here as the system output. As shown earlier in Eq. (81), its
260 SHIMA, IDAN, AND GOLAN

relative degree is one. Because in this analysis the model uncertain- All of the terms in the numerator of Eq. (A21) were shown to be
ties and disturbances are zero, Eq. (81) simplifies to bounded. Similarly to the SMG proof of the preceding section, be-
cause in the endgame Vr < 0, the fin control command δ c is bounded
Ż GC = {Vλ + aTN τT [1 − exp(−tgo /τT )] + ψα (tgo )α + ψq (tgo )q except possibly for tgo → 0.
  Using Eqs. (4), (A3b), and (A15c), we obtain that for r → 0, while
+ ψδ (tgo )δ}V̇r r Vr2 + ψδ (tgo )δ c τs (A13) on the sliding surface,
The differential equations for the first five states in Eq. (A12), while
2V̇r λ̇τs
on the sliding surface are given by Eqs. (A3). The nonlinear dy- δ c (tgo → 0) ≈ (A22)
namics equations for α and q are given by Eqs. (8), (39), and (40). L̄ δ cos(γ M0 − λ0 )
The canard deflection angle, while on the sliding surface, can be
obtained from Eq. (33): Boundedness of λ̇ and V̇r imply also boundedness of the fin de-
 flection command. This completes the proof that all of the internal
δ = Vr tgo
2
λ̇ − aTN τT2 ψ(tgo /τT ) − ψα (tgo )α − ψq (tgo )q ψδ (tgo ) states of the integrated guidance-control problem and the fin deflec-
tion command, when solved using the integrated SMGC approach,
(A14) are bounded when on the sliding surface Z GC (t) ≡ 0.
Note that the first seven states are not directly affected by the system
input δ c . Therefore, the system dynamics for the new states are References
expressed in the normal form, where [r λ aT γT γ M α q]T are 1 Blakelock, J. H., Automatic Control of Aircraft and Missiles, 2nd ed.,
its internal states.3
Downloaded by SIMON FRASER UNIVERSITY on October 9, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.14951

Wiley, New York, 1991, Chap. 7.


We now examine the boundedness of the internal states in our 2 Lin, C. F., Wang, Q., Speyer, J. L., Evers, J. H., and Cloutier, J. R.,
finite horizon problem, assuming we are on the sliding surface “Integrated Estimation, Guidance, and Control System Design Using Game
Z GC = 0. From Eqs. (8), (39), (40), and (41), we obtain that, be- Theoretic Approach,” Proceedings of the American Control Conference,
cause f i (·), i = 1, 2, 3, 4 are bounded functions, the internal missile American Automatic Control Council, Evanston, IL, 1992, pp. 3220–3224.
3 Khalil, H. K., Nonlinear Systems, 3rd ed., Prentice–Hall, Upper Saddle
dynamics states α, q, and the missile acceleration a M are bounded.
The latter implies also boundedness of γ M . Based on the arguments River, NJ, 2002, Chap. 13.
4 Menon, P. K., and Ohlmeyer, E. J., “Nonlinear Integrated Guidance-
provided in the preceding subsection, the internal states r , aT , and
Control Laws for Homing Missiles,” AIAA Paper 2001-4160, 2001.
γT are bounded. In addition, λ̇ is bounded, except maybe for r → 0. 5 Menon, P. K., Sweriduk, G. D., and Ohlmeyer, E. J., “Optimal Fixed-
From Eq. (A14), assuming ψδ (ν) < 0 ∀ ν > 0 (this assumption has Interval Integrated Guidance-Control Laws for Hit-to-Kill Missiles,” AIAA
been verified numerically), we obtain that on the surface the fin Paper 2003-5579, 2003.
deflection angle δ is also bounded, except maybe for r → 0. 6 Palumbo, N. F., and Jackson, T. D., “Integrated Missile Guidance and
Using a series expansion of GC from Eq. (34), we obtain Control: A State Dependent Riccati Differential Equation Approach,” Pro-
 ceedings of the IEEE International Conference on Control Applications,
ψα (tgo → 0) ≈ − L̄ α cos(γ M0 − λ0 )tgo
2
2 (A15a) Vol. 1, IEEE Press, Piscataway, NJ, 1999, pp. 243–248.
7 Utkin, V. I., Sliding Modes in Control and Optimization, Springer-Verlag,
 Berlin, 1992.
ψq (tgo → 0) ≈ − L̄ α cos(γ M0 − λ0 )tgo 6
3
(A15b) 8 Shkolnikov, I. A., and Shtessel, Y. B., “Aircraft Nonminimum Phase
 Control in Dynamic Sliding Manifolds,” Journal of Guidance, Control, and
ψδ (tgo → 0) ≈ − L̄ δ cos(γ M0 − λ0 )tgo 2
2
(A15c) Dynamics, Vol. 24, No. 3, 2001, pp. 566–572.
9 Slotine, J.-J. E., and Li, W., Applied Nonlinear Control, Prentice–Hall,
Using Eq. (A14), we obtain that for r → 0 (or equivalently tgo → 0) Upper Saddle River, NJ, 1991, Chap. 7, pp. 276–307.
and while on the sliding surface 10 Salamci, M. U., Özgören, M. K., and Banks, S. P., “Sliding Mode Con-
trol with Optimal Sliding Surfaces for Missile Autopilot Design,” Journal
δ(tgo → 0) ≈ [−2Vr Vλ /r + aTN − L̄ α α cos(γ M0 − λ0 )]/ of Guidance, Control, and Dynamics, Vol. 23, No. 4, 2000, pp. 719–727.
11 Shkolnikov, I., Shtessel, Y. B., Lianos, D., and Thies, A. T., “Robust
[ L̄ δ cos(γ M0 − λ0 )] (A16) Missile Autopilot Design via High-Order Sliding Mode Control,” AIAA
Substituting this result into Eq. (64b), we obtain that Paper 2000-3968, 2000.
12 Thukral, A., and Innocenti, M., “A Sliding Mode Missile Pitch Autopilot

V̇λ (tgo → 0) ≈ Vr Vλ /r + aM (A17) Synthesis for High Angle of Attack Maneuvering,” IEEE Transactions on
Control Systems Technology, Vol. 6, No. 3, 1998, pp. 359–371.
13 Moon, J., and Kim, Y., “Design of Missile Guidance Law Via Variable
where
Structure Control,” Journal of Guidance, Control, and Dynamics, Vol. 24,
aM = ã M cos(γ M0 − λ0 ) − a M cos(γ M − λ) (A18) No. 6, 2001, pp. 659–664.
14 Xu, W., Mu, C., and Zhou, D., “Adaptive Sliding-Mode Guidance of
and a Homing Missile,” Journal of Guidance, Control, and Dynamics, Vol. 22,
No. 4, 1999, pp. 589–594.
ã M = L̄ α α + L̄ δ δ (A19) 15 Shkolnikov, I., Shtessel, Y. B., and Lianos, D., “Integrated Guidance-

Because a M , α, and δ are bounded, then so is aM . Using Eq. (A17) Control System of a Homing Interceptor: Sliding Mode Approach,” AIAA
in Eq. (64c) leads to Paper 2001-4218, 2001.
16 Zarchan, P., Tactical and Strategic Missile Guidance, Progress in Astro-

λ̈(tgo → 0) ≈ aM (A20) nautics and Aeronautics, Vol. 176, AIAA, Washington, DC, 1997, pp. 27–28.
17 Shima, T., and Shinar, J., “Time Varying Pursuit Evasion Game Mod-

Because we are dealing with a finite-time interception problem, λ̇ els with Bounded Controls,” Journal of Guidance, Control, and Dynamics,
and consequently the internal state λ are bounded. Vol. 25, No. 3, 2002, pp. 425–432.
18 Gutman, S., “On Optimal Guidance for Homing Missiles,” Journal of
The commanded fin deflection angle, while on the sliding surface,
Guidance, Control, and Dynamics, Vol. 3, No. 4, 1979, pp. 296–300.
is obtained by setting Eq. (A13) to zero, resulting in 19 Shinar, J., Solution Techniques for Realistic Pursuit-Evasion Games,
  Advances in Control and Dynamic Systems, Vol. 17, Academic Press,
δ c = −V̇r r τs Vr2 ψδ (tgo ) {Vλ + aTN τT [1 − exp(−tgo /τT )] New York, 1981, pp. 63–124.
20 Friedland, B., Control System Design: An Introduction to State Space
+ ψα (tgo )α + ψq (tgo )q + ψδ (tgo )δ} (A21) Methods, McGraw–Hill, New York, 1986, pp. 247, 248.

You might also like