You are on page 1of 14

Proceedings of the 18th World Congress

The International Federation of Automatic Control


Milano (Italy) August 28 - September 2, 2011

A Generic Approach to Missile Autopilot Design


using State-Dependent Nonlinear Control
Tayfun Çimen*

* ROKETSAN Missiles Industries Inc., Ankara, TURKEY


(e-mail: tcimen.control@gmail.com, tcimen@roketsan.com.tr).

Abstract: The fundamental objective of autopilot design for missile systems is to provide stability with
satisfactory performance and robustness over the whole range of flight conditions throughout the entire
flight envelope that missiles are required to operate in, during all probable engagements. Depending on the
control mode (skid-to-turn or bank-to-turn), intercept scenario (such as surface to surface, surface to air,
air to air) and mission phase (launch, midcourse, terminal), missile autopilots can command accelerations,
body rates, incidence angles, or flight path angles. To this end, classical and modern multivariable
techniques from linear control theory combined with gain scheduling have dominated missile autopilot
design over the past several decades. In this paper, the concept of extended linearization (also known as
state-dependent coefficient parameterization) is examined for state-dependent nonlinear formulation of
the vehicle dynamics in a novel and very general form for the development of a generic and practical
autopilot design approach for missile flight control systems. Any extended linearization control method,
such as the currently popular State-Dependent Riccati Equation (SDRE) methods, can then be applied to
this state-dependent formulation for missile flight control system design. The unique contribution of this
paper is the novel use of a very general and realistic nonlinear aerodynamic model that captures all major
aerodynamic nonlinearities attributed to missiles, together with the fully nonlinear and coupled 6-DOF
equations of motion of rigid-body missile dynamics for full-envelope, 3-axes nonlinear autopilot design,
without invoking any of the usual simplifying assumptions of the traditional linear design philosophy, and
independent of any flight or trim conditions. Moreover, in the development of the generic approach, all the
autopilot command structures mentioned above are incorporated in one compact topology. Practical
considerations such as actuator dynamics and actuator position and rate saturation are also included in the
development of the nonlinear autopilot. The proposed approach has been implemented and its performance
and robustness validated in detailed 6-DOF simulations in three dimensional environments, using various
missile configurations with stable, unstable and nonminimum-phase characteristics.
Keywords: Actuator saturation; Aerospace applications; Autopilot design; Flight control systems; Missile
dynamics; Nonlinear control; Nonlinear systems; Optimal control; Riccati equations.

1. INTRODUCTION missile dynamics, neglecting all aerodynamic cross couplings


between channels. Clearly, an autopilot derived from
Missiles are required to operate over a large flight envelope to
linearization about a single flight condition will be unable to
meet the challenge of highly maneuverable targets in all
achieve suitable performance over all envisioned operating
probable engagements, such as surface-to-air and air-to-air
conditions. Thus, standard practice in designing missile
missions against threats posed by both tactical aircraft and
autopilots is to linearize the equations of motion for various
missiles. The fundamental objective of autopilot design for
flight conditions (around certain finite number of operating
missile systems is to provide stability with satisfactory
points and trim conditions in the flight envelope), and apply
performance and robustness over the whole range of flight
linear design techniques to these static models to deliver the
conditions that may be encountered throughout the flight
desired performance characteristics in the local region of the
envelope. However, the equations of motion that govern the
fixed operating point. Satisfactory performance across the
behavior of a flight vehicle in six (three translational and three
flight envelope may then be achieved by gain scheduling the
rotational) degrees of freedom (DOF) are both nonlinear and
resulting sets of local autopilot controller parameters as
time varying, making flight control system design for missiles
functions of the flight conditions (such as Mach number,
a complex, nontrivial multi-input–multi-output control
altitude and angle of attack) to yield a global controller.
problem. In view of the uncertainties in aerodynamic
parameters, cross-coupling effects, nonlinearities, and Although gain scheduling is the most prominent approach in
measurement inaccuracies, the task of guaranteeing stability, missile autopilot design, there are some drawbacks to this
performance and robustness throughout the entire flight design philosophy. The first is that due to wide variations in
envelope is a challenging one. flight conditions, such as those encountered by high angle-of-
Conventional missile autopilot design is based on the attack missiles, linear autopilots require gain scheduling with
application of linear control techniques to the decoupled and respect to Mach number, altitude, angle of attack and sideslip
linearized longitudinal and lateral equations of motion of angle. Therefore, depending on the mission and the flight

978-3-902661-93-7/11/$20.00 © 2011 IFAC 9587 10.3182/20110828-6-IT-1002.03744


18th IFAC World Congress (IFAC'11)
Milano (Italy) August 28 - September 2, 2011

envelope, conventional autopilots with gain scheduling may the design does not involve any decoupling of flight
necessitate controller design for several flight conditions. dynamics, does not require gain scheduling, is independent of
Although powerful classical and multivariable linear trim conditions, and is very straightforward and practical.
techniques are available for design, gain scheduling is a very
The paper is organized as follows. For completeness, Section
tedious process that can consume an enormous amount of
2 first provides an overview of the general form of the 6-DOF
time and effort. Secondly, control design methods based on
rigid-body equations of motion for flight vehicle dynamics,
linear time-invariant models ignore the fundamental
clearly stating any assumptions, definitions and conventions
(nonlinear) nature of the plant for flight control systems. The
used in deriving the model. These equations are required in
controller so designed may offer unsatisfactory performance
deriving the nonlinear dynamic equations for state-dependent
and are severely limited, especially when the dynamics is
nonlinear formulation of the dynamics for missile autopilot
highly nonlinear and undergoes large motion. The ever
design, and to illustrate how the 3-axes, coupled dynamics,
increasing high performance and greater maneuverability
which are valid at high angle of attack and sideslip angle,
requirements demand that missiles operate in regimes of large
differ from conventional decoupled dynamics. The derivation
angles and angular rates, where nonlinearities are dominant
of this nonlinear design model is presented in Section 3, along
and a high degree of coupling exists between the lateral and
with actuator characteristics. For the purposes of this study,
longitudinal motions because of inertial coupling. Under these
only aerodynamic control surfaces (canard, wing or tail) are
circumstances, the assumption of small motion about an
considered as missile control effectors, leaving thrust vector
operating point does not hold true and nonlinear approaches
control and reaction jet control thrusters as straightforward
must be employed for autopilot design.
extensions of the developed approach. The proposed nonlinear
In comparison to linear designs, nonlinear autopilots require autopilot topology, presented in Section 4, exploits the time-
significantly higher initial analysis effort, but are considerably scale separation between the long-period (phugoid) and short-
easier to design than gain-scheduled linear controllers. Due to period modes, as well as between “slow” translational
current technological trends and the sufficient computational dynamics and “fast” rotational dynamics of the short-period
resources that are nowadays available on-board, nonlinear mode inherent in missiles and other flight vehicles. In Section
autopilots have become viable candidates for implementation 5, the dynamics derived in Section 3 are formulated in state-
in missiles. Owing to these reasons and the severe nonlinear dependent nonlinear form. In Section 6, a fully coupled,
characteristics of flight vehicle dynamics at large angles and nonlinear approach is taken to design a high-performance, 3-
angular rates, there is obvious temptation for control designers axes, full-envelope missile autopilot based on the state-
to venture outside the linear design domain, and select dependent nonlinear formulation of Section 5. In order to
nonlinear design methods. validate the satisfactory performance, robustness, genericity
and practicality of the nonlinear missile autopilot design
In this paper, a generic and practical approach to nonlinear
approach developed in this study, detailed 6-DOF nonlinear
missile autopilot design is developed through a novel state-
simulations have been carried out in three dimensional
dependent nonlinear formulation of missile aerodynamics and
environments, under several noteworthy, realistic and coercive
6-DOF rigid-body dynamics, using the concept of extended
engagements, with various missile configurations. However,
linearization, also known as state-dependent coefficient
due to space limitations, only a brief discussion is pursued
(SDC) parameterization (Çimen, 2010). The development is
about the configurations tested, simulations performed and the
pursued independent of any flight (or trim) condition and
results obtained, leaving discussions on design details,
without invoking any of the usual simplifying assumptions in
engagement scenarios, performance and robustness
missile autopilot design, such as constant mass (post burnout),
evaluations, and missile flight control system simulations for
constant speed or constant air density, and without neglecting
these test cases for presentation during the Congress. The
any cross coupling effects, such as roll rate, sideslip, and yaw
paper is then closed with a brief summary in Section 7.
rate. The design is based on a much more general and realistic
missile model compared to the plethora of nonlinear 2. MISSILE FLIGHT DYNAMICS
approaches previously considered in the literature.
2.1 Assumptions, Definitions and Conventions
In the development of the nonlinear missile autopilot, all
nonlinearities of missile dynamics are taken into full account. The first objective is to derive the general form of equations
Specifically, the model captures all major nonlinear variations of motion for a flight vehicle in 6 DOF (for further reading,
of the aerodynamic coefficients with flight parameters, and readers may refer to Blakelock, 1991; Etkin, 2005; Stevens &
includes the inertial coupling between roll, pitch and yaw Lewis, 2003; Zipfel, 2007). The derivation is carried out
channels as well as the asymmetric structure of the missile under the following assumptions:
airframe. The genericity to the approach is achieved through A1) The effects of structural deflections (aeroelasticity) and
the use of a very general and realistic nonlinear aerodynamic of the dynamics of the relative motion of the control
model together with the fully coupled, nonlinear model for the surfaces on overall missile dynamics are neglected,
6-DOF rigid-body dynamics, by developing a high- such that missile dynamics is based on the equations of
performance, full-envelope, 3-axes nonlinear missile autopilot motion for a single rigid body.
design concept. Through this genericity, the approach A2) The rotation of the Earth is neglected, with the curved
becomes applicable to both skid-to-turn (STT) and bank-to- surface (longitude/latitude grid) of the fixed Earth
turn (BTT) missiles (whether canard, wing or tail controlled), unwrapped into a flat plane tangential to the launch
and during all phases of flight. More importantly, however, point. This is the so called stationary and flat-Earth

9588
18th IFAC World Congress (IFAC'11)
Milano (Italy) August 28 - September 2, 2011

approach applicable to in-atmosphere, Earth-bound frame FV (also known as the flight path frame). Since air is at
flights, valid for vehicles flying in the atmosphere with rest by Assumption A3, the two velocity vectors are in fact
speeds less than Mach 5 (below hypersonic speed). one and the same, so that the X V coordinate axis also lies
A3) The air is assumed to be at rest, neglecting any effect along the total velocity vector V of the vehicle CM, as
caused by wind. shown in Fig. 1. Two angles γ and χ then describe the
A4) Gravity is assumed to be acting through the center of instantaneous angular orientation of the velocity coordinates
mass of the missile, so that missile center of mass relative to the Earth coordinate system. The horizontal flight
(CM) coincides with missile center of gravity (CG). path angle (or heading angle) χ is measured from North to
Note that these are standard assumptions for deriving the 6- the projection of V into the local tangent plane X E YE
DOF rigid-body equations of motion of a flight vehicle, and (positive clockwise about downward vertical Z E ), such that
all are applicable to tactical missiles and, with the exception North and East are expressed by 0 and π 2 rad, respectively.
of Assumption A3, to aircraft as well. The vertical flight path angle γ takes this projection
vertically up to V . Note that only X W of wind coordinates
For modeling the motion of a rigid-body flight vehicle under has been defined without ambiguity, and so its orientation is
Assumption A1, two reference frames are introduced. They now defined relative to velocity coordinates through the bank
are the fixed-Earth frame FE (OE ; i E , jE , k E ) and the missile (roll) angle μ . The wind and velocity coordinates are then
body-fixed frame FB (OB ; i B , jB , k B ) , whose triad of unit base related through μ about the velocity vector, such that the
vectors are orthonormal and right-handed by convention. The velocity frame FV is obtained by rotating FW about the X W
axes of the coordinate systems associated with each of the axis so that the YV axis of FV becomes parallel to the
frames FE and FB are aligned with their respective triads, as horizontal plane of FE .
shown in Fig 1. Due to assumption A2, the fixed-Earth frame
FE is declared as an inertial (Newtonian) reference frame. Its
origin OE is defined as the launch point of the missile, with
X E -axis pointing North, YE -axis pointing East, and the
positive Z E -axis pointing downward in the direction of the
local gravity vector. The body-fixed frame FB has its origin
OB located at the instantaneous CM of the missile, the X B -
axis is parallel with the body longitudinal centerline and
points toward the nose of the missile, the YB -axis is directed
to the right when viewed from the rear, and the Z B -axis
points in a direction below the horizontal. The three Euler
angles φ , θ and ψ in roll, pitch and yaw, respectively, then
describe the instantaneous angular orientation of the body
coordinate system relative to the Earth coordinate system.
In the sequel, 6-DOF rigid-body nonlinear equations of
motion will be derived, and then resolved in the body
reference frame, such that all components of forces and
moments acting on the missile will be modeled in the missile
body-fixed axes. Fig. 1 shows the conventions used for
positive directions of the vector components of force,
moment, linear (translational) velocity, and angular
(rotational) velocity of a missile resolved in the body
coordinate system. The nomenclature adopted here is Fig. 1: Fixed-Earth and body-fixed reference frames showing
presented in Table 1. The six projections of the linear and vector components resolved in Cartesian body coordinates
angular velocity vectors on the moving body coordinate axes
are the six degrees of freedom. Table 1: Nomenclature for vector components resolved in Fb
As the vehicle moves through the air mass, it experiences a Vector Roll Axis Pitch Axis Yaw Axis
Description Units
OB X B OBYB OB Z B
relative wind over its body, which gives rise to aerodynamic Symbol
FX B FYB FZ B
forces. Introducing the wind frame FW , such that its X W F Total Force N
coordinate axis lies along the velocity vector V of the vehicle M Total Moment Nm L M N
CM w.r.t. the air mass (see Fig. 1), gives rise to the Cartesian V Linear Velocity ms −1 u v w
incidence angles (also called wind angles), which are the ωB E Angular Velocity rads −1 p q r
pitch-plane angle of attack α and yaw-plane sideslip angle β. a Total Acceleration ms −2 aX B aYB aZ B

The wind frame relates “the velocity vector of the CM w.r.t.


the air mass” to the body frame. In order to relate the “ground 2.2 Coordinate Transformations
(inertial) velocity vector of the CM w.r.t. the Earth” (that is, Fig. 2 summarizes the different coordinate systems defined in
velocity vector of the CM w.r.t. the air mass + wind velocity Section 2.1, and depicts the transformation angles that relate
w.r.t. inertial fixed-Earth frame) to the inertial fixed-Earth them to each other. Now, let us present the orthonormal
frame, an additional frame is introduced, called the velocity transformations that connect all these coordinate systems.

9589
18th IFAC World Congress (IFAC'11)
Milano (Italy) August 28 - September 2, 2011

FE
Note that the mass is not constant in a missile during the
operation of the propulsion system because of the
φ , θ ,ψ γ,χ consumption of propellant. It can be shown that (5) is
applicable to a missile that has variable mass due to the
burning of propellant if the value of m to be used in (5) is a
α,β μ
FB FW FV function of time t . A derivation that rigorously takes into
account the change in mass, and arrives at the same result for
Fig. 2: Coordinate systems and transformation angles constant mass, except that m is now a function of time, is
given in Meriam & Kraige (1986).
The Euler angles φ , θ , and ψ relate the body coordinates to
the Earth coordinates. The coordinate (or component) The vector differential equation (5), although valid in any
transformation matrix of body w.r.t. inertial coordinates is reference frame, is simple enough to be expressed in the body
obtained using a 3-2-1 (yaw-pitch-roll) Euler sequence, giving frame FB by resolving it into Cartesian body coordinates.
⎡1 0 0 ⎤ ⎡ cθ 0 − sθ ⎤ ⎡ cψ sψ 0⎤ Thus, using the vector components of V , ω B E and F in FB
⎢ ⎥ (1)
T( B , E ) = ⎢0 cφ sφ ⎥ ⎢⎢ 0 1 0 ⎥⎥ ⎢⎢ − sψ cψ 0 ⎥⎥ , defined in Table 1, yields the scalar equations for d B V ( B ) dt :
⎢0 − sφ cφ ⎥⎦ ⎢⎣ sθ 0 cθ ⎥⎦ ⎢⎣ 0 1 ⎥⎦
⎣ 0 u = rv − qw + FX B m ⎫
with s (⋅) and c(⋅) denoting shorthand notations for sin(⋅) and ⎪⎪
v = pw − ru + FYB m ⎬ (6)
cos(⋅) , respectively. ⎪
w = qu − pv + FZ B m .⎪⎭
The orientation of body coordinates relative to wind The three first-order differential equations (6) govern the
coordinates is related through the angle of attack α and dynamics of the 3-DOF translational motions of a vehicle
sideslip angle β . By noting that the wind coordinates are a expressed in the body frame with the Earth as the inertial
transformation of the body coordinates using negative β and reference frame. These equations are nonlinear and coupled
positive α rotation sequence, the transformation matrix by the body rates ω(BBE) = [ p q r ]T . The nonlinearity enters
T( B ,W ) of body coordinates w.r.t. wind coordinates becomes through incidence angles in aerodynamic force and moment
⎡ cα 0 − sα ⎤ ⎡ c β − sβ 0⎤
(2) calculations (discussed later in Sections 2.5, 2.6 and 3.1).
T( B ,W ) = ⎢⎢ 0 1 0 ⎥⎥ ⎢⎢ s β cβ 0 ⎥⎥ .
⎣⎢ sα 0 cα ⎦⎥ ⎣⎢ 0 0 1 ⎦⎥
2.4 Rotational Motion
The orientation of wind coordinates relative to velocity
coordinates is related through the bank angle μ . The The three rotational degrees of freedom are governed by
transformation matrix T(W ,V ) of wind coordinates w.r.t. Euler’s law. The rotational equations of motion of a flight
velocity coordinates is given by vehicle w.r.t. a nonrotating and flat Earth (Assumption A2)
⎡1 0 0⎤ are expressed by the vector differential equation
ω B E = I −1 (t ) ( M − ω B E × I (t )ω B E ) ,
⎢ ⎥ (3)
T(W ,V ) = ⎢0 c μ sμ ⎥ . (7)
⎢0 − s μ c μ ⎥⎦
⎣ where I(t ) is the (variable) inertia matrix of missile body
Finally, the sequence of rotations χ and γ relate velocity referred to the CM, Iω B E is the angular momentum of body
coordinates to Earth coordinates, giving w.r.t. the Earth frame referred to the CM, and M represents
⎡ cγ 0 − sγ ⎤ ⎡ c χ sχ 0⎤ the vector sum of all the external moments referred to the
⎢ ⎥ (4)
T(V , E ) = ⎢ 0 1 0 ⎥ ⎢⎢ − s χ cχ 0 ⎥⎥ . missile CM. Although the vector differential equation (7) is
⎢ sγ 0 cγ ⎥⎦ ⎢⎣ 0 0 1 ⎥⎦
⎣ also valid in any reference frame, the body coordinates
express the inertia matrix I in constant form, defined as
2.3 Translational Motion ⎡ Ix − I xy − I xz ⎤
⎢ ⎥ (8)
I ( B ) = ⎢ − I xy Iy − I yz ⎥ ,
The three translational degrees of freedom are governed by ⎢ − I xz
⎣ − I yz I z ⎥⎦
Newton’s second law. Frequently, the trajectory of a vehicle
is so slow compared to orbital speed and so close to the Earth which is a symmetric and positive-definite (nonsingular)
that, under the nonrotating (stationary) and flat Earth matrix whose elements consist of the moments of inertia
Assumption A2, the Coriolis and centrifugal accelerations I x , I y , I z about the axes of FB , and the products of inertia
related to the rotation of the Earth and the changes in the I xy , I yz , I xz over the planes of FB , all having SI units of
direction of gravity as the missile moves over the surface of kg m 2 . Since the reference frame is fixed to the body, the
the Earth can be neglected. Then, under Assumption A2, the inertia matrix appears constant in FB , and does not change by
translational equations of motion of a flight vehicle over a body motion. Note, however, that during the operation of the
nonrotating and flat Earth take on the vector form missile propulsion system, there is another source of change
d E V dt = d B V dt + ω B E × V = F m(t ) , (5) in the inertia matrix that is not related to body motion. As the
propellant mass is expelled from the missile, the inertia matrix
where m(t ) is the vehicle (variable) mass, V is the linear changes. Therefore, similar to the variable mass m , the
velocity vector of the missile CM w.r.t. to the Earth, d E V dt parameter I is not constant in a missile during the operation
and d B V dt are its Earth-frame and body-frame derivatives, of the propulsion system because of the consumption of
respectively, ω B E is the angular velocity of FB w.r.t. to FE , propellant. Although a rigorous accounting for a variable
and F represents the vector sum of external forces applied inertia matrix is not easy, the time rate of change of I due to
directly to the missile, such that the acceleration of the missile the burning of propellant is usually small enough that terms
CM w.r.t. to the Earth is given by a = d E V dt = F m . expressing this rate can be neglected. By this assumption, (7)

9590
18th IFAC World Congress (IFAC'11)
Milano (Italy) August 28 - September 2, 2011

is also applicable to a missile that has variable inertia matrix, where V ( B ) = [u v w]T , V (W ) = [V 0 0]T , and V is the
again provided that the value of I to be used in (7) is a airspeed, given by the magnitude of the missile velocity
function of time t . vector, such that
V 2 = u 2 + v 2 + w2 . (11)
Often, because of symmetry, missiles have body axes that
Using (2) and (10), the components of the total velocity vector
coincide with the principal axes of the moment of inertias, and
in body coordinates are thus obtained as
the inertia matrix exhibits only diagonal elements, which
u = V cα c β , v = V s β , w = V sα c β . (12)
greatly simplifies the equations of motion, since the products-
of-inertia terms vanish. Moreover, the mass distribution of a The definitions for the Cartesian incidence angles are derived
missile about its YB -axis is often essentially the same as that from (12) in terms of the velocity components in body
about its Z B -axis. This is a particularly important case for coordinates ( u , v , w ) and the relative wind axis component
conventional missiles having tetragonal symmetry that skid to (airspeed V ), given by the algebraic relationships
turn, and manifest two equal principal moment of inertias. α = tan −1 wu , β = sin −1 Vv . (13)
Therefore, the further simplification of setting I yy = I zz is
often possible; however, the distinction is retained here for 2.6 Forces and Moments
generality. On the other hand, bank-to-turn or cruise missiles
Equations (6) and (9) express the translational and rotational
have aircraft-like symmetries, where all of the principal
equations of motion in body coordinates. With the exception
moment of inertia axes are not aligned with the body axes.
of Assumptions A1-A4, these equations are perfectly general,
However, due to the planar symmetry w.r.t. the X B and Z B
that is, no simplifying assumptions have been used in their
axes (that is, I xy = I yz = 0 ), the YB axis is indeed a principal
derivation. Solution of these equations of motion requires
axis. Such a configuration gives rise to an inertia matrix in
knowledge of the sum of the forces and the sum of the
body coordinates with diagonal elements consisting of axial
moments acting on the missile. The forces and moments are
moments of inertia I xx and I zz , and the principal moment of
produced by gravity, aerodynamics, and propulsion. Since, by
inertia I yy , and one off-diagonal element I xz , leading to more
Assumption A4, missile CM and CG are the same point, the
complex solutions. Since the product-of-inertia term I xz , in
gravitational force does not cause any external moment about
general, is nonzero, it is retained in the development of the
the CM. Consequently, gravity contributes only as external
rotational equations of motion.
force acting on the missile body axes. Therefore, the forces
For missiles with planar symmetry only in the X B Z B plane, and moments are generally composed of:
there are four zero elements in the inertia matrix (8), so that • gravitational forces,
its inversion can be tolerated, making it possible to solve for • aerodynamic forces and moments, and
the body rate components ω(BBE) = [ p q r ]T explicitly as • propulsive thrust forces and any moment caused either
scalar equations. The rates of change of the body rate by design (such as thrust vectoring and/or reaction jet
components are thus obtained in scalar form by expressing (7) controls) or by error (misalignment of the thrust).
in the body frame FB , giving The total force applied to the missile body, therefore, consists
p= 1
I x I z − I xz 2
{I L + I
z N + [ I z ( I y − I z ) − I xz 2 ]qr + I xz ( I x − I y + I z ) pq} ⎫
xz

of gravitational, aerodynamic and propulsive forces Fg , Fa
⎪ (9) and Fp , respectively, whereas the total moment acting on the
q = I1y {M + ( I z − I x ) pr − I xz ( p 2 − r 2 )} ⎬ body generally consists of aerodynamic and propulsive

r = I I − I 2 { I xz L + I x N + [ I x ( I x − I y ) + I xz ] pq + I xz ( I y − I z − I x )qr}.⎪
1 2 moments M a and M p , respectively, giving F = Fg + Fa + Fp
x z xz ⎭ and M = M a + M p .
The three first-order differential equations (9) govern the
dynamics of the 3-DOF rotational motions of a vehicle w.r.t. In aerodynamically controlled missiles, propulsive thrust is
the inertial frame expressed in body axes. These equations are usually designed to act through the missile CM, and thus
nonlinear and couple with the translational equations (6) only produces no moment about the CM (that is, M p = 0 ). When
through the aerodynamic moments, which are contained in the the thrust vector does not pass through the CM, either by
total moments M ( B ) = [ L M N ]T . Their integration yields the design or error, the equations used to describe the resulting
angular velocity of the body w.r.t. the inertial frame expressed propulsive moment M p on the missile is equal to the product
in body coordinates, in short, the body rates ω(BBE) = [ p q r ]T . of the magnitude of the thrust and the perpendicular distance
between the thrust vector and the CM. In this study, only the
Clearly, the product of inertia I xz couples the pitch and yaw generic design approach for aerodynamically controlled
rates to the roll degrees of freedom, and a strong coupling missiles will be developed in order to highlight the steps
exists for large values of I xz . The pitch and yaw equations involved in the missile autopilot design procedure, without
exhibit similar trends. Other phenomena caused by I xz are the further complicating the nonlinear equations of motion.
coupling of the yawing moment N into the roll axis and the However, the formulation presented here can readily be
rolling moment L into the yaw axis. Therefore, if I xz = 0 , as extended to combining thrust vectoring and reaction jets into
is the case for missiles with tetragonal symmetry, the missiles’ aerodynamic control system.
rotational equations of motion (9) are greatly simplified.
The flat-Earth model is represented by having the
2.5 Incidence Angles gravitational force always pointing along the Z E axis in the
inertial coordinate system, such that g ( E ) is formulated in the
The transformation of the missile velocity vector from relative form [0 0 g ]T , where g is the acceleration due to gravity.
wind coordinates to body coordinates is obtained from Hence, gravity in body coordinates is modeled as
V ( B ) = T( B ,W ) V (W ) , (10) g ( B ) = T( B , E ) g ( E ) , using the transformation matrix T( B , E ) given

9591
18th IFAC World Congress (IFAC'11)
Milano (Italy) August 28 - September 2, 2011

in (1). The gravitational force acting on the body can then be 3.1 Aerodynamic Model
obtained from Fg( B ) = mg ( B ) . The components of gravitational
In this work, a generic and realistic aerodynamic model,
force resolved into body coordinates are thus obtained as
applicable as a general rule to missiles, is proposed. Because
Fg( B ) = mg[− sθ sφ cθ cφ cθ ]T . (14)
of the complicated functional dependence of the aerodynamic
Aerodynamic forces and moments acting on the body are coefficients represented in (19), in practice, each “total”
expressed in body coordinates using the nondimensional coefficient is modeled as the sum of a baseline component
aerodynamic (axial, side and normal) body-axes force (that are, individually, functions of fewer variables), plus
coefficients C X (−C A ) , CY , CZ (−CN ) , nondimensional incremental terms. The baseline components are primarily
aerodynamic (rolling, pitching and yawing) body-axes functions of M , α , β , δ a , δ e , δ r , whereas incremental terms
moment coefficients ClCG , CmCG , CnCG acting about the CG, are functions of only M . Component buildups of the “total”
the dynamic pressure q , missile reference (characteristic) nondimensional aerodynamic coefficients for missiles, in
length d (such as maximum diameter), and missile reference general, are then expressed in the form
area S (such as maximum cross-sectional area), as follows: C X = C X′ j ( M , h, α , β , δ a , δ e , δ r ) for j = thrust on/off ⎫
Fa( B ) = qS [C X CY CZ ]T , M (aB ) = qSd [Cl Cm Cn ]T . (15) ⎪
CG CG CG CY = CY′ ( M , α , β , δ r ) + 2dV [CYr ( M )r + CYβ ( M ) β + CYpα ( M ) pα ] ⎪
⎪ (20)
The dynamic pressure q is a function of airspeed V and air CZ = CZ′ ( M , α , β , δ e ) + 2V [CZ q ( M )q + CZα ( M )α + CYpα ( M ) p β ]
d
⎪⎪
density ρ (which, in turn, is a function of the altitude h ), ClCG = Cl′RP ( M , α , β , δ a , δ e , δ r ) + 2dV Cl p ( M ) p


and is obtained from ⎪
CmCG = Cm′ RP ( M , α , β , δ e ) − lCZ + 2dV [Cmq ( M )q + Cmα ( M )α + Cn pα ( M ) p β ]⎪
q = 1 2 ρV 2 . (16)

CnCG = Cn′RP ( M , α , β , δ r ) + lCY + 2dV [Cnr ( M )r + Cnβ ( M ) β + Cn pα ( M ) pα ], ⎪
Since, by assumption, the thrust vector Fp passes along the ⎭
X B axis and through the missile CM, Fp does not create any where all angular-rate variables are nondimensionalized. The
external moment, giving baseline components are the first terms on the right-hand side
Fp( B ) = [T 0 0]T , M (pB ) = [0 0 0]T . (17) of the equations, which are followed by incremental terms, in
The total forces and moments acting on the body can now be the order of importance, consisting of the angular-rate
obtained from (14)-(17), which are expressed in component variables. The coefficients of these incremental terms,
form in body coordinates as indicated by an additional level of subscripts, represent
aerodynamic derivatives, which are partial derivatives of the
⎡ FX ⎤ ⎡ −mg sθ + qSC X + T ⎤ ⎡ L ⎤ ⎡ qSdClCG ⎤
⎢ B⎥ ⎢ ⎥ ⎢ ⎥ (18) aerodynamic force and moment coefficients, all having
F ( B ) = ⎢ FYB ⎥ = ⎢ mg sφ cθ + qSCY ⎥ , M ( B ) = ⎢⎢ M ⎥⎥ = ⎢ qSdCmCG ⎥ .
functional dependence on the Mach number M . Note that the
⎢ ⎥ ⎢ mg c c + qSC ⎥ ⎢⎣ N ⎥⎦ ⎢⎢ qSdC ⎥⎥
⎣⎢ FZ B ⎦⎥ ⎣ φ θ Z ⎦
⎣ nCG ⎦ Magnus terms CYpα and Cn pα in (20) are particularly
Aerodynamic forces and moments are modeled directly in the significant for agile missiles with canard-controlled and
body axes, using body-axes aerodynamic force and moment rolling airframe configurations, in which case they should be
coefficients. These coefficients are typically functions of included in the autopilot model (Nielsen, 1988).
flight conditions, such as Mach number ( M ), altitude ( h ),
incidence angles ( α , β ) and their rates ( α , β ), body angular Due to the significant changes in CG position during
rates ( p , q , r ), effective (virtual) aerodynamic control surface propellant consumption, aerodynamic rolling, pitching and
deflections ( δ a , δ e , δ r ), changes in CG, and whether the yawing moments are generally measured about a fixed
thrust power is on or off (referred to as plume effects). Thus, reference point (RP) at a distance of xRP meters from the
missile nose (in the case the reference point is the missile
C j = C j ( M , h, α , β , α ∗ , β ∗ , p∗ , q∗ , r ∗ , δ a , δ e , δ r , CG, thrust on/off), (19) nose, then xRP = 0 ), thus obtaining ClRP , CmRP and CnRP . In
j = X , Y , Z , lCG , mCG , nCG , (⋅)∗ = d
2V
(⋅), this way, the implicit dependence of the aerodynamic moment
with (⋅)∗ representing nondimensionalized motion related coefficients on the varying CG position is avoided. The
variables for angular rates. These complicated and highly moment coefficients about the CG ( ClCG , CmCG and CnCG ) are
nonlinear functions are used in the equations of motion to then obtained by shifting moments about the fixed RP to the
model the airframe’s aerodynamics. Note that, since all instantaneous CG position, located xCG meters away from
coefficients are unitless, all angles are in radians. Note also missile nose, using the moment arm xCG − xRP . Although this
that the control surface deflections can be canard, wing or tail. is assumed to have no effect on ClCG ( ClCG = ClRP ), the shifting
of aerodynamic moments from the RP to the CG couples CZ
3. DERIVATION OF THE NONLINEAR DYNAMIC into CmCG and CY into CnCG , as is evident from the latter two
EQUATIONS FOR MISSILE AUTOPILOT DESIGN expressions in (20), where l ( xCG − xRP ) d . In doing so, the
dependence of these coefficients on the instantaneous CG
With the general equations of motion of a flight vehicle over a position, which varies during propellant consumption, is
flat Earth in place, we can now proceed with the first step explicitly expressed.
necessary for autopilot development. This will be deriving the
equations of motion required for nonlinear missile autopilot Data for baseline components and incremental terms in (20)
design, which is pursued for all variables used in the are derived from empirical methods, aerodynamic prediction
aerodynamic coefficients. Since aerodynamic coefficients for programs, computational modeling, and/or through wind-
missiles are generally modeled as having functional tunnel experiments, and are compiled in the aerodynamic
dependence on several variables given in (19), the equations database generally in the form of look-up tables. Incremental
of motion for autopilot design will be developed based on at terms consist of 1-dimensional look-up tables, since these are
least these variables, in addition to the command variables functions of only the Mach number. However, look-up tables
that will be conveyed to the autopilot. for baseline components are necessarily multi-dimensional.

9592
18th IFAC World Congress (IFAC'11)
Milano (Italy) August 28 - September 2, 2011

For the development of the autopilot model, general nonlinear symmetric w.r.t. the origin. However, these terms will not be
representations of baseline components will now be presented. zero for unsymmetrical data, in which case they must be
The proposed model uses nonlinear representations up to retained. In addition, inclusion of second-order nonlinearities
cubic terms for baseline components of “total” aerodynamic α α and β β in the expressions provides more accurate
force and moment coefficients, and captures the most estimates to those obtained using α 3 and β 3 terms alone,
significant nonlinearities attributed to missiles, as follows: respectively, and vice versa. Some of the neglected terms
C X′ j = C X 0 + a1α 2 + a2 β 2 + a3αδ e + a4 βδ r + a5δ e 2 + a6δ r 2 + a7δ a 2 ⎫ (such as higher order terms), on the other hand, are often
⎪ negligible, but can be included in the expressions if desired.
CY′ = CY0 + CYδ δ r + b1 β + b2 β 2 + b3 β β + b4 β 3 + b5α 2 β ⎪
r

CZ′ = CZ0 + CZδ δ e + c1α + c2α + c3 α α + c4α + c5αβ
2 3 2
⎪⎪ (21)
e
3.2 Mach Dynamics


ClRP = Cl0 + d1δ a + d 2α δ a + d 3 β δ a + d 4αδ e + d5 βδ r
2 2


Aerodynamic coefficients are modeled as having functional
Cm′ RP = Cm0 + Cmδ δ e + e1α + e2α 2 + e3 α α + e4α 3 + e5αβ 2 ⎪ dependence on the Mach number M , as opposed to body-
e
⎪ axes components u , v , w or the airspeed V . A differential
Cn′RP = Cn0 + Cnδ δ r + f1 β + f 2 β 2 + f 3 β β + f 4 β 3 + f 5α 2 β , ⎪⎭
equation describing the dynamics of M is thus required,
r

where the newly introduced aerodynamic derivatives (such as which can be derived from the algebraic relationship
CYδ , CZδ , etc.) as well as the rest of the coefficients (such as
r e M =V a, (22)
a1 , b1 , etc.) in these expressions all have functional
dependence on only the Mach number M , although not where a is the speed of sound. Assuming a standard
explicitly shown in (21) for notational convenience. atmospheric model, the speed of sound is a function of the
altitude h . Thus, differentiating (22) gives
The proposed form of the aerodynamic coefficients given in M = Vaa−2Va .
(21) can be obtained using data fitting techniques. Noting that
there is linearity with respect to the unknown coefficients of However, inside the troposphere (the lowest portion of the
the baseline components in (21), linear least squares method Earth’s atmosphere, which extends from sea level up to 11
can be used to estimate these parameters. The best fit to the km), the speed of sound changes only around 10% in value.
tabulated aerodynamic data using the proposed model (21) is Therefore, even though h and a are both varying, the time
then obtained by solving the resulting overdetermined system. rate of change of a is so small that, in the derivation of the
For example, suppose that the baseline component CZ′ is differential equation for Mach number, it can be assumed
represented in tabular form as a function of only the three zero, giving
variables M , α and δ e . Hence, from (21), M ≅ V a. (23)
CZ′ = CZ0 ( M ) + CZδ ( M )δ e + c1 ( M )α + c2 (M )α 2 + c3 ( M ) α α + c4 ( M )α 3 , The equation for V is obtained by differentiating (11), and
e

which is a multinomial in two independent variables α and substituting (6) to give


δ e . Suppose also that CZ′ is tabulated, say, using x , y and V = V1 (uu + vv + ww)
z breakpoints of M , α and δ e , respectively. This results in = V1 [u ( a X B − qw + rv) + v(aYB − ru + pw) + w(aZ B − pv + qu )],
yz equations for 6 unknowns for each set of Mach
where a X B = FX B m , aYB = FYB m , and aZ B = FZ B m are the
breakpoints k = 1, … , x , and yields an overdetermined system
components of total acceleration in body coordinates resulting
A k x k = b k , where A k ∈ yz ×6 , b k ∈ yz×1 and
from the gravitational, aerodynamic and thrust forces acting
xk = [CZ0 , k CZδ c1, k c2,k c3, k c4, k ]T
e
,k
on the vehicle. Substituting (12) into this equation yields the
with yz 6 . The best fit that gives the optimal parameter differential equation for airspeed, that is,
values xˆ k is then found by solving the corresponding normal V = a X cα c β + aY s β + aZ sα c β .
B B B
(24)
equation of the overdetermined system, that is,
Hence, substituting (24) into (23) yields
( A k T A k )xˆ k = A k T b k . This is carried out for the complete set
M = 1a (a X B cα c β + aYB s β + aZ B sα c β ). (25)
of Mach breakpoints k = 1, … , x . Then, if desired, the 6
coefficients solved at each Mach breakpoint can be obtained
in analytical form as functions of Mach number, by fitting an 3.3 Incidence Angle Dynamics
appropriate function (such as polynomial, Fourier, etc.) to Using body coordinates, the translational equations of motion
these data. Another option is to store these coefficients in 1- (6) are solved for the rates of change of the components u , v
dimensional look-up tables and use interpolation techniques to and w of the total velocity vector V . Their integration yields
obtain intermediate values online. u , v and w , which are used to calculate α and β using
Nonlinear and coupled terms in (21) represent the dominant (13). The use of the wind coordinate system causes the rate of
nonlinearities in the baseline components of the missile change of the magnitude of the velocity vector ( V ) to also lie
aerodynamic force and moment coefficients. Coupled terms in along the X W wind axis, and yields explicit equations for the
these expressions produce coupling between channels at high rates of change of α and β . The following derivation will
angle of attack. Depending on the application, some of these form a set of differential equations describing the dynamics of
terms may have negligible effect on missile aerodynamics α and β , valid for large α and β < 90° .
and, hence, may be neglected. For example, some of the terms The equations for α and β are obtained by differentiating
at zero incidence angles and control deflections ( CY0 , CZ0 , (13) and using (12) to yield
Cl0 , Cm0 , Cn0 ) will be zero, except for asymmetric airframes/ w cα − u sα v −V s β
airfoils. Similarly, α 2 and β 2 terms in the expressions for α= uw − wu
u 2 + w2
= V cβ
, β= Vv − vV
= V cβ
.
V V 2 − v2
CY′ , CZ′ , Cm′ and Cn′ will be obtained as zero when data is Substituting (6), (22) and (24) into these equations yields

9593
18th IFAC World Congress (IFAC'11)
Milano (Italy) August 28 - September 2, 2011

α = q − ( p cα + r sα ) t β + Ma1cβ (aZ cα − a X sα ) B B
(26) is, position and rate limits of the aerodynamic control
surfaces, as shown in Fig. 4. Actuator dynamics are, therefore,
β = p sα − r cα − Ma
1
( a X cα s β − aY c β + aZ sα s β ),
B B B
(27) represented by the 2nd-order nonlinear differential equation
where t (⋅) denotes tan(⋅) . δ i = (δ icom − sat(δ i , δ max ))ωn 2 − 2ζωn 2 sat(δ i , δ max ) (29)
using the saturation function
3.4 Actuator Dynamics ⎧σ max , σ > σ max

Missiles operate using four aerodynamic control surfaces, sat(σ , σ max ) ⎨σ , σ ≤ σ max
⎪−σ , σ < −σ max
using either canards or wings or tail fins. Generally, these ⎩ max
four control surfaces actuate independently, each driven by a with σ = δ i , δ i , for i = 1, … , 4 .
separate but identical actuator (servomotor). These actuators rate limit position limit
δi δi δi
are required for tracking the guidance commands (such as δ icom +

ωn 2 +
− ∫
δ max

δ max
sat(δ i , δ max )

lateral and normal acceleration commands expressed in g’s)


sat(δ i , δ max )
through actual control surface deflections δ i via commanded 2ζωn
control surface deflections δ icom , with i = 1, … , 4 . Therefore,
the effective (virtual) control surface deflection commands of Fig. 4. 2nd-Order Nonlinear Actuator Block Diagram Model
aileron (roll), elevator (pitch), and rudder (yaw) control Therefore, defining the actual and the commanded control
( δ a , δ e , δ r ) from the missile autopilot are distributed to the surface deflection vectors, respectively, as
four (actual) individual control surface deflections δ i δ [δ1 δ 2 δ 3 δ 4 ]T
( i = 1, … , 4 ). The control mixing logic that relates these
δ com [δ1com δ 2com δ 3com δ 4com ]T ,
effective control surface deflections to individual control
surface deflections is configuration specific, and depends (29) can be represented in state-space form as
upon whether the missile flies with a “+” or “×” aerodynamic ⎡δ ⎤ ⎡ 0 4× 4 I 4×4 ⎤ ⎡δ ⎤ ⎡ 0 ⎤
⎢ ⎥=⎢ ⎥ ⎢ ⎥ + ⎢ 24×4 ⎥ δ com (30)
control (or “delta”) configuration, as shown in Fig. 3. The ⎣δ ⎦ ⎢⎣ −ωn diag
2
δi {
sat(δ i ,δ max )
} −2ζωn diag 2
{ sat(δ i ,δ max )
δi } ⎥⎦ ⎣δ ⎦ ⎣ωn I 4×4 ⎦
arrows in this figure show the direction of forces and leading
for i = 1, … , 4 , where 04× 4 and I 4× 4 are 4 × 4 zero and
edge of panel where, by convention, a positive panel
identity matrices, respectively, and diag{} ⋅ represents a
deflection is one which will produce a negative
diagonal matrix.
(counterclockwise when viewed from the rear) rolling
moment increment at zero angle of attack and sideslip. The 3.5 Body Rate Dynamics
equations for the control surface mixing logic are then given by
[δ a δ e δ r ]T = Δ logic [δ1 δ 2 δ 3 δ 4 ]T . (28) Expanding the equations derived in (9), substituting the total
moments (18) and aerodynamic coefficients (20)-(21), and
Two conventional mixing logics for the “+” and “×” delta
using the derivations obtained for α and β in (26) and (27),
configurations, respectively, are defined by
respectively, yields the following nonlinear representations
⎡1 1 1 1 ⎤ ⎡1 1 1 1 ⎤
+
Δ logic =
1⎢
0 2 0 −2 ⎥⎥ , Δ ×logic =
1⎢
1 1 −1 −1⎥⎥ .
for missile body angular rate dynamics:
4⎢ 4⎢ p = Lbias + L p p + Lr r + L pq pq + Lqr qr + Lδ a δ a + Lδ e δ e + Lδ r δ r ⎫
⎢⎣ 2 0 −2 0 ⎥⎦ ⎢⎣1 −1 −1 1 ⎥⎦
⎪⎪
q = M bias + M p p + M q q + M r r + M p2 p + M r 2 r + M pr pr + M δ e δ e ⎬ 2 2 (31)
(+) (–) (–) (+) ⎪
4 1 r = N bias + N p p + N r r + N pq pq + N qr qr + Nδ a δ a + Nδ e δ e + Nδ r δ r ⎪⎭
1
(–) (+)
(+) (–) where the coefficients are defined as
4 2 negative rolling Lbias kI z Cl0 + kI xz [(Cn0 + lCY0 ) + ( f1 + lb1 ) β + ( f 2 + lb2 ) β 2 + ( f3 + lb3 ) β β
moment
+ ( f 4 + lb4 ) β 3 + ( f5 + lb5 )α 2 β − 2dV 1
Ma (Cnβ + lCYβ )( a X B cα s β − aYB c β + aZ B sα s β )]
(–) (+)
(+) (–) Lp d
2V k[ I z Cl p + I xz (Cn pα + lCYpα )α + I xz (Cnβ + lCYβ ) sα ]
3
(–) (+)
3 2 Lr d
2V kI xz [(Cnr + lCYr ) − (Cnβ + lCYβ ) cα ] L pq I xz ( I x − I y + I z ) ( I x I z − I xz 2 )
(+) (–)
Lqr [ I z ( I y − I z ) − I xz ] ( I x I z − I xz )
2 2
Lδ a kI z (d1 + d 2α 2 + d3 β 2 )
Fig. 3. Rear-View Projection of “+”, “×” Delta Configurations
Lδ e kI z d 4α Lδ r k[ I z d5 β + I xz (Cnδ + lCYδ )]
For satisfactory autopilot performance, actuator dynamics
r r

must also be considered. It is usually sufficient to consider a M bias qSd


Iy [(Cm0 − lCZ0 ) + (e1 − lc1 )α + (e2 − lc2 )α + (e3 − lc3 ) α α + (e4 − lc4 )α
2 3

second-order actuator dynamic model, with transfer function + (e5 − lc5 )αβ 2 + 2dV 1
Ma cβ (Cmα − lCZα )(aZ B cα − a X B sα )]
δi ( s ) ωn 2
δ icom ( s )
= s 2 + 2ζωn s + ωn 2
, i = 1, … , 4, Mp qSd 2
2VI y [(Cnpα − lCYpα ) β − (Cmα − lCZα ) cα t β ]
qSd 2 2
Mq [(Cmq − lCZ q ) + (Cmα − lCZα )] Mr − qSd
2VI y (Cmα − lCZα ) sα t β
where ζ is the actuator damping ratio, ωn is the natural 2VI y

frequency, δ icom is the commanded control surface deflection M p2 − I xz I y M r2 I xz I y


of the i th aerodynamic control surface, and δ i is its actual M pr (I z − I x ) I y M δe qSd
Iy (Cmδ − lCZδ )
deflection. However, actuators possess significant
e e

nonlinearities, such as angular position and rate limits as well N bias kI xz Cl0 + kI x [(Cn0 + lCY0 ) + ( f1 + lb1 ) β + ( f 2 + lb2 ) β 2 + ( f3 + lb3 ) β β
as mechanical backlash, and it is preferable to include these + ( f 4 + lb4 ) β 3 + ( f5 + lb5 )α 2 β − 2dV 1
Ma (Cnβ + lCYβ )(a X B cα s β − aYB c β + aZ B sα s β )]
nonlinearities in the actuator model. It is important to note Np d
2V k[ I xz Cl p + I x (Cn pα + lCYpα )α + I x (Cnβ + lCYβ ) sα ]
that it is δ i that exhibit these nonlinearities (not δ a , δ e , δ r ). Nr d
2V kI x [(Cnr + lCYr ) − (Cnβ + lCYβ ) cα ] N pq [ I x ( I x − I y ) + I xz 2 ] ( I x I z − I xz 2 )
In the design and the analysis of the nonlinear autopilot, N qr I xz ( I y − I z − I x ) ( I x I z − I xz 2 ) Nδ a kI xz [d1 + d 2α 2 + d3 β 2 ]
actuator dynamics are modeled using actuator saturation, that Nδ e kI xz d 4α Nδ r k[ I xz d5 β + I x (Cnδ + lCYδ )]
r r

9594
18th IFAC World Congress (IFAC'11)
Milano (Italy) August 28 - September 2, 2011

which are highly nonlinear functions of M , α , β , α , β , V , Y′ 2CY′ − ( Mad )2 CYβ (a X B cα s β − aYB c β + aZ B sα s β ) Z′ 2CZ′ + ( Mad)2 c CZα (aZ B cα − a X B sα )
β

with k qSd ( I x I z − I xz 2 ) , q = 1 2 ρV 2 and l = ( xcg − xref ) d . YaX λ (Y ′ cα c β − CYα sα c β − CYβ cα s β ) Z aX λ ( Z ′ cα c β − CZα sα c β − CZ β cα s β )


YaY λ (Y ′ s β + CYβ c β ) Z aY λ ( Z ′ s β + CZ β c β )
Although not explicitly shown, recall that the moment and
YaZ λ (Y ′ sα c β + CYα cα c β − CYβ sα s β ) Z aZ λ ( Z ′ sα c β + CZα cα c β − CZβ sα s β )
product of inertia terms in these expressions are time-varying
YM 2 λ aCYδ δ r ZM 2 λ aCZδ δ e
due to changes in CG position during propellant consumption. r

λ M 2 a (−CYα cα t β + CYβ sα )
e

Yp Zp λ M 2 a(−CZα cα t β + CZβ sα )
+λ Md (CYpα α + CYβ sα ) + g cφ cθ + λ Md (CYpα β − CZα cα t β ) − g sφ cθ
3.6 Acceleration Dynamics
Yq λ M 2 aCYα Zq λ M 2 aCZα + λ Md (CZ + CZα ) − g sθ
q

First, a relation between the Euler angular rates φ , θ , ψ and Yr −λ M 2 a (CYα sα t β + CYβ cα ) Zr −λ M 2 a (CZα sα t β + CZβ cα )
the body-fixed angular rates ω(BBE) = [ p q r ]T is required. +λ Md (CYr − CYβ cα ) + g sθ −λ MdCZα sα t β

Using a ψ - θ - φ rotation sequence from fixed-Earth with λ 2 m ρ aS . Here C X


1 ′ , CY′ , CZ′ are obtained directly from
coordinates to body-fixed coordinates and then inverting this (21), whereas C X and M from (20) and (25), respectively.
matrix equation yields
⎡ p ⎤ ⎡1 0 − sθ ⎤ ⎡ φ ⎤ ⎡ φ ⎤ ⎡ 1 sφ t θ cφ tθ ⎤ ⎡ p ⎤ 3.7 Bank Angle Dynamics
⎢ q ⎥ = ⎢0 c ⎥⎢ ⎥ ⎢ ⎥ ⎢ ⎥
⎢ ⎥ ⎢ sφ cθ ⎥ ⎢ θ ⎥ ⇒ ⎢θ ⎥ = ⎢0 cφ − sφ ⎥ ⎢⎢ q ⎥⎥ .
φ
⎢⎣ r ⎥⎦ ⎢⎣ 0 − sφ cφ cθ ⎥⎦ ⎢⎣ψ ⎥⎦ ⎢ψ ⎥ ⎢ 0 sφ cθ cφ cθ ⎥⎦ ⎢⎣ r ⎥⎦
The respective angular velocities ωW E and ω B E of the wind
⎣ ⎦ ⎣
frame FW and the body frame FB w.r.t. the Earth frame FE
Let us now derive the rigid-body acceleration dynamics for are related by
the missile body acceleration components at the CM. The
ω B E = ωW − β k W + α jB .
components of acceleration in body coordinates, E

a( B ) = [a X B aYB aZ B ]T , are obtained from the components of Resolving this equation in body coordinates gives
total force acting at the CM along the missile body coordinate ω(BBE) = T( B ,W ) ωW(W E) − β T( B ,W ) k W + α jB ,
axes, F ( B ) = [ FX B FYB FZ B ]T , divided by the mass m . Now, so that ωW(W E) can be expressed in component form as
taking into account the dependence of q on V in (16), and ωW(W E) = T (W , B ) [ p − β sα q −α r + β cα ]T , (33)
recalling the functional dependence of the aerodynamic force (W , B )
coefficients on flight conditions α , β and the aerodynamic with T obtained from (2). A relationship between the
control surfaces δ a , δ e , δ r , differentiating a( B ) = F ( B ) m angular rates μ , γ , χ and the wind-fixed angular rates ωW(W E)
using (18), keeping terms up to first order, ignoring the time can also be derived using a χ - γ - μ sequence of rotations
rate of change of thrust, substituting for the Euler angular from fixed-Earth coordinates to wind coordinates, giving
⎡1 0 − sγ ⎤ ⎡ μ ⎤
rates, and grouping terms yields the acceleration dynamics ⎢ ⎥
ωW(W E) = ⎢ 0 c μ s μ cγ ⎥ ⎢⎢ γ ⎥⎥ .
⎡aX B ⎤ ⎛ ⎡ 2V C C Xα C X β ⎤ ⎡V ⎤ ⎡C X δa C Xδ C X δ ⎤ ⎡δ ⎤ ⎞
⎡ 0 − cφ cθ sφ cθ ⎤ ⎡ p ⎤ ⎢0 − s μ
⎢ ⎥ ⎢ ⎥
⎜⎢ X
⎥⎢ ⎥ ⎢ e r
⎥⎢ a ⎥⎟
⎣ c μ cγ ⎥⎦ ⎢⎣ χ ⎥⎦
⎢ aYB ⎥ = g ⎢ cφ cθ 0 sθ ⎥ ⎢⎢ q ⎥⎥ + m1 qS ⎜ ⎢ 2 V CY CYα CYβ ⎥ ⎢α ⎥ + ⎢ CYδ CYδ CYδ ⎥ ⎢δ e ⎥ ⎟ .
⎜⎢ ⎢ a ⎥⎢ ⎥⎟
⎢ ⎥ ⎥ Inverting this matrix equation and substituting into (33) gives
e r

a
⎣⎢ Z B ⎦⎥
⎢ − sφ cθ
⎣ − sθ 0 ⎥⎦ ⎣⎢ r ⎦⎥ ⎜ ⎢ 2 V CZ CZα CZ β ⎦⎥ ⎣⎢ β ⎦⎥ ⎢ CZδ C Zδ C Zδ ⎥ ⎣ δ r ⎦ ⎟
⎝⎣ ⎣ a e r ⎦ ⎠
⎡ μ ⎤ ⎡1 t γ s μ t γ c μ ⎤ ⎡ cα c β sβ sα c β ⎤ ⎡ p − β sα ⎤
Substituting (22), (24), (26) and (27) into this equation, the ⎢ γ ⎥ = ⎢0 cμ

− s μ ⎥ ⎢⎢ − cα sβ cβ
⎢ ⎥
− sα sβ ⎥⎥ ⎢ q − α ⎥ .
⎢ ⎥ ⎢
acceleration dynamics can be represented in the compact form ⎢ ⎥
⎣⎢ χ ⎦⎥ ⎢⎣ 0 s μ cγ ⎥
c μ cγ ⎦ ⎣⎢ − sα 0 cα ⎦⎥ ⎣ r + β cα ⎦
⎡aX B ⎤ ⎡ 2 Ma C X C Xα C X β ⎤ ⎡ Ma c c Ma s β Ma sα c β ⎤ ⎡ a X B ⎤

⎢ aYB ⎥ =
⎥ 1
qS

⎢ 2 Ma CY CYα
⎥⎢ α β

CYβ ⎥ ⎢ − sα c β 0
⎥⎢
cα c β ⎥ ⎢ aYB ⎥
⎥ The required equation for μ is then obtained from the above
relation by expanding it and substituting (26) and (27) for α
mMa
⎢ ⎥ ⎢ ⎥⎢
CZ β ⎦⎥ ⎣ − cα s β cβ − sα s β ⎥⎦ ⎢⎢ aZ ⎥⎥
⎣⎢ aZ B ⎦⎥ ⎣⎢ Ma CZ C Zα
2
⎣ B⎦
⎡ 2 Ma C X C X C X ⎤ ⎡ 0 and β , respectively, thus obtaining the bank angle dynamics
0 0 ⎤ ⎡ p⎤ ⎡ 0 − cφ cθ sφ cθ ⎤ ⎡ p ⎤
⎢ α β

+ m1 qS ⎢ 2 Ma CY CYα CYβ ⎥ ⎢⎢ − cα t β

1 − sα t β ⎥⎥ ⎢⎢ q ⎥⎥ + g ⎢ cφ cθ 0

sθ ⎥ ⎢⎢ q ⎥⎥ μ = p ccαβ + r csαβ + Ma
1
[a X (sα t β + sα t γ s μ − cα s β t γ c μ )
⎢ ⎥ ⎢ 0 ⎦⎥ ⎣⎢ r ⎦⎥
B
(34)
⎢⎣ 2 Ma CZ CZα CZ β ⎥⎦ ⎣⎢ sα 0 − cα ⎦⎥ ⎣⎢ r ⎦⎥ ⎣ − sφ cθ − sθ
+ aYB (c β t γ c μ ) − aZ B (sα s β t γ c μ + cα t β + cα t γ s μ )].
⎡C X C X δ C X δ ⎤ ⎡δ ⎤
⎢ δa e r
⎥⎢ a⎥
+ m1 qS ⎢ CYδ CYδ CYδ ⎥ ⎢δ e ⎥ .
⎢ a e r
⎥⎢ ⎥
⎢⎣ CZδa CZδe CZδr ⎥⎦ ⎣δ r ⎦
3.8 Flight Path Angle Dynamics
Therefore, expanding these equations, substituting the For a flat Earth,
aerodynamic force coefficients from (20)-(21), taking partial d E V dt = dV V dt + ωV E × V = F m(t ) ,
derivatives of (20)-(21) to obtain aerodynamic derivatives of which can be expressed in velocity coordinates as
force coefficients (that is, C X α = ∂C X ∂α , etc.), and noting dV V (V ) dt + ωV(V E) × V (V ) = F(V ) m , (35)
that CYδ = CYδ = CZδ = CZδ = 0 , algebraic manipulations
a e a r where V = [V 0 0] and thus dV V dt = [V 0 0]T .
(V ) T (V )
yield the following nonlinear representations of missile body
The angular velocity ωV E of the velocity frame FV w.r.t. the
acceleration dynamics:
Earth frame FE consists of the vector addition of the flight
a X B = ( X aX a X B + X aY aYB + X aZ aZ B ) M + X M 2 M 2 + X p p + X q q + X r r ⎫
⎪⎪ path angular rates χ and γ times their respective unit vectors
aYB = (Ya X a X B + YaY aYB + YaZ aZ B ) M + YM 2 M 2 + Yp p + Yq q + Yr r ⎬ (32) so that, by resolving ωV E in velocity coordinates, the

aZ B = ( Z aX a X B + Z aY aYB + Z aZ aZ B ) M + Z M 2 M 2 + Z p p + Z q q + Z r r , ⎪⎭ following coordinated equation is obtained for expressing the
where velocity-fixed angular rates ωV(V/)E in component form:
C Xα = 2a1α + a3δ e CYα ≈ 2b5αβ + 2dV CYpα p ⎡0 ⎤ ⎡ cγ 0 − s γ ⎤ ⎡ 0 ⎤ ⎡ − χ sγ ⎤
CZα ≈ c1 + 2c2α + c3 α + 3c4α 2 + c5 β 2 ⎢ ⎥ ⎢ ⎥
C X β = 2a2 β + a4δ r CYβ ≈ b1 + 2b2 β + b3 β + 3b4 β 2 + b5α 2 CZβ ≈ 2c5αβ + 2dV CYpα p ωV(V E) = γ ⎢⎢1 ⎥⎥ + χ ⎢ 0 1 0 ⎥ ⎢⎢ 0 ⎥⎥ = ⎢ γ ⎥ .
C X δ = 2a7δ a CXδ = a3α + 2a5δ e CXδ = a4 β + 2a6δ r ⎢⎣ 0 ⎥⎦ ⎢ sγ 0 cγ ⎥⎦ ⎢⎣1 ⎥⎦ ⎢⎣ χ cγ ⎥⎦
a
e r

λ (2C X′ cα c β − C Xα sα c β − C X β cα s β )
Now, since F (V ) = Fg (V ) + Fa (V ) + Fp (V ) ,
X aX Xp λ M 2 a(−C Xα cα t β + C X β sα )
F(V ) = mT(V , E ) g( E ) + T(V ,W ) T(W , B ) [Fa ( B ) + Fp ( B ) ].
j

X aY λ (2C X′ s β + C X β c β )
j
Xq λ M 2 aC Xα − g cφ cθ
X aZ λ (2C X′ sα c β + C Xα cα c β − C X β sα s β )
j
Xr −λ M 2 a (C Xα sα t β + C X β cα ) + g sφ cθ Noting that Fa ( B ) + Fp ( B ) = m[a ( B ) − g ( B ) ] yields
XM2 λ a(C Xδ δ a + C X δ δ e + C X δ δ r )
a e r F (V ) m = T(V , E ) g ( E ) + T(V ,W ) T(W , B ) [a( B ) − g ( B ) ],

9595
18th IFAC World Congress (IFAC'11)
Milano (Italy) August 28 - September 2, 2011

which is now conveniently expressed in terms of the body for missile flight control systems. In the development of this
accelerations a( B ) . Substituting these expressions for ωV(V E) generic approach, the proposed topology will accommodate
and F (V ) m into (35) with (22) for V , using the coordinate all the abovementioned autopilot structures.
transformation matrices derived in Section 2.2, and solving
for γ and χ , the expressions for flight path angle dynamics 4.2 Autopilot Design Topology
are obtained as In flight dynamics, short-period mode is primarily described
γ= 1
[(sα c μ + cα s β s μ )a X B − c β s μ aYB − (cα c μ − sα sβ s μ )aZ B + g1 ] ⎫⎪ by body rates and incidence angles, whereas long-period
Ma

(36)
χ= 1
Ma cγ
[(sα s μ − cα s β c μ )a X B + c β c μ aYB − (cα s μ + sα s β c μ )aZ B + g 2 ], ⎪

(phugoid) mode is primarily described by flight path angles.
In addition to this two-layer structure that exists between
where
short-period and long-period modes, flight vehicles also
g1 g[(cα s β sθ + c β sφ cθ − sα sβ cφ cθ )s μ + (sα sθ + cα cφ cθ ) cμ − cγ ]
experience inherent time-scale separation between “slow”
g2 − g[(cα s β sθ + c β sφ cθ − sα s β cφ cθ ) cμ + (sα sθ + cα cφ cθ ) sμ ]. translational dynamics and “fast” rotational dynamics of the
short-period mode. This phenomenon causes autopilot design
3.9 Altitude Dynamics philosophies based on a single unified (single loop)
framework to be ineffective, because control surface
Transforming the components of V (V ) to the fixed-Earth deflections directly respond to the translational error
coordinate system gives correction demands, which may lead to instability of the
V ( E ) = T( E ,V ) V (V ) = V [cγ c χ cγ s χ − sγ ]T . rotational dynamics. This is especially true for control
Thus the dynamics for the altitude h is obtained from the surfaces located either at the front or the tail of the missile,
negative Z E component of V ( E ) , so that using (22) gives because deflections of these control surfaces can create only
h = Ma sγ . (37) minor forces, whereas they create large moments due to a
long moment arm from the CM. Consequently, these control
4. AUTOPILOT STRUCTURE surfaces are ineffective in directly correcting translational
errors, whereas they can be very effective in turning the flight
All equations are now in place to proceed with nonlinear vehicle. Therefore, for a successful flight control system, the
autopilot development, which is based on the dynamic models design must exploit the time-scale separation that exists
derived in Section 3. Note that these dynamic models are very between translational and rotational motions of the CM.
general compared to all other nonlinear approaches proposed
in the literature for missile autopilot design. In the development of the generic approach to missile
autopilot design in this study, the proposed structure explicitly
4.1 Autopilot Design Models exploits the inherent time-scale separation that exists between
translational and rotational motions of flight vehicles, as well
Maximizing overall missile performance requires choosing as between short-period and long-period motions. The former
the appropriate autopilot command structure for each mission issue is addressed using a two-loop autopilot design topology,
phase. This may include designing different autopilots for as opposed to using a single-loop structure. The latter issue is
launch, an agile turn at high angle of attack, midcourse where addressed by an additional outer loop, giving the three-loop
a long flyout is required, and endgame where terminal homing structure presented in Fig. 5. The outer loop converts flight
maneuvers are necessary. The missile autopilot can command path angle commands ( γ com , χ com ) or altitude commands
accelerations, body rates, incidence angles, and flight path ( h com ) from the guidance system to bank angle commands,
angles. For example, vertical launch surface-air missiles may together with either incidence angle commands ( α com , β com )
require flight path angle commands during the launch phase. or lateral-normal (pitch and yaw) body acceleration
For air-air missiles, on the other hand, a body rate command commands ( aYcom , aYcom ) for the intermediate loop. The
is typically used during launch, since rate-command autopilots B B
intermediate loop then converts these incidence angles or
are very robust to the uncertain proximity aerodynamics body accelerations to body roll rate, pitch rate and yaw rate
(Wise & Broy, 1998). An agile turn requires directional commands ( p com , q com , r com ) for the inner loop. Finally, the
control of the missile’s velocity vector relative to the missile inner loop converts p com , q com , r com to actual control surface
body (given by incidence angles). This necessitates following deflection commands δ icom ( i = 1, … , 4 ) for the actuators.
angle-of-attack and sideslip angle commands, and regulating
roll to zero. Finally, during midcourse and terminal phases, an In the design, full nonlinear and coupled 6-DOF equations of
acceleration autopilot is generally used. Some missiles, such motion describing missile dynamics derived in Section 3 are
as cruise missiles, also employ altitude-hold autopilots. directly used in all three loops, without making any
simplifications to these equations. This is achieved by
In what follows, a generic approach for nonlinear missile formulating the equations in state-dependent nonlinear form
autopilot design will be developed using the complete set of for each loop for controller design, which is discussed next.
nonlinear dynamic equations (20)-(37) derived in Section 3
Outer-Loop Intermediate-Loop Inner-Loop Missile
"γ com , χ com " + + + Actuators
− Controller μ com − Controller p com − Controller δ icom δi Dynamics
or " hcom " qcom
"α com , β com "
or " aYcom , aZcom " r com
B B

μ p, q, r , δ i , δ i
"α , β " p, q, r
a X B , aYB , aZ B , M , α , β , μ Sensors
or " aYB , aZ B "
"γ , χ " &
γ , χ, h
or " h " Observers

Fig. 5. Three-Loop Autopilot Design Topology

9596
18th IFAC World Congress (IFAC'11)
Milano (Italy) August 28 - September 2, 2011

5. STATE-DEPENDENT NONLINEAR FORMULATION 5.1 State-Dependent Formulation of the Inner Loop


The first objective is to represent the equations for each loop Consider the body-rate dynamics derived in (31), and note the
in Fig. 5 in the input-affine form presence of the state-independent terms (referred to as bias
x(t ) = f (x) + B(x)u(t ), x(0) = x0 (38) terms) Lbias , M bias , N bias , which prevents a direct C1
with state vector x ∈ n and (unconstrained) input vector factorization of f (x) into A(x, α )x for SDC representation of
u ∈ m , where f (x) is a continuously differentiable vector- (31). In general, within the framework of state-dependent
valued function of x , that is, f (⋅) ∈ C1 , and the origin x = 0 parameterization, bias terms are handled by multiplying and
is an equilibrium point of the system with u = 0 , such that dividing these terms by a state that will never go to zero
f (0) = 0 , and B(x) ≠ 0 ∀x . Readers may refer to Çimen (Cloutier & Stansbery, 2001; Çimen, 2010). However, the
(2010) for an overview of the various methods available for three states p , q , r can all go to zero. Alternatively, if the
systematically handling numerous systems that do not meet bias term is constant or slowly-varying, then it can be
the basic structure and conditions mentioned above. Once the modeled as a stable state b(t ) = −λ b(t ) , where λ is a small
systems meet the proper structure and conditions, the concept positive number. Each time through the controller, the actual
of extended linearization is used to formulate the nonlinear value of b(t ) would then be used in calculating the control u .
vehicle dynamics in SDC form (Çimen, 2010). Thus, the However, Lbias , M bias and N bias are neither constant nor
input-affine nonlinear system (38) is represented in the form slowly-varying due to their dependence on M , α , β , α , β
x(t ) = A(x)x(t ) + B(x)u(t ), x(0) = x0 , (39) and q . A third alternative is to introduce an additional state
which has a linear structure with SDC matrices A(x) and s with stable dynamics s (t ) = −λs s (t ) , λs > 0 , for the
B(x) . Note that in a deterministic setting, the SDC purpose of absorbing the biases. Augmenting the system with
parameterization (39) fully captures the nonlinearities of the the stable state, bias terms can then be factored as b = [b s ]s ,
system. Although the SDC parameterization is unique in the where b denotes the biases Lbias , M bias , N bias . Each time
case of scalar x for all x , it is nonunique in the through the inner-loop controller, the initial value s (0)
multivariable case, and the SDC parameterization A(x) itself (assumed small) is used in the SDC matrix in calculating u .
can be parameterized as A(x, σ ) , where σ is a vector of free
Thus, defining the state vector x [ p q r s ]T and control
design parameters (see Proposition 1 in Çimen, 2010). The
vector u [δ a δ e δ r ]T , (31) can now be represented
introduction of σ creates additional degrees of freedom that
(nonuniquely) in SDC form (39). Following the systematic
can be used to enhance controller performance, avoid
procedure described above, any term containing more than
singularities or loss of controllability, and affect tradeoffs
between optimality, stability, robustness, and disturbance one state variable must first be parameterized and then
rejection, thus offering a very flexible nonlinear control policy. apportioned among the corresponding elements of the A
matrix. There are five such terms in (31), each containing a
The primary condition in selecting the “right” A(x) is that pair of state variables. Thus, introducing the vector σ1 with
the respective pairs {A (x), B(x)} are pointwise stabilizable five design parameters σ 1,1 , … , σ 1,5 ∈ , these terms can all
SDC parameterizations of the nonlinear system (38) ∀x ∈ Ω . be parameterized, and a family of SDC parameterizations for
In addition to satisfying this requirement, a rule of thumb in the system can then be constructed with
selecting the state-dependent factorization is that any term ⎡ a11 a12 a13 1
Lbias ⎤ ⎡ Lδ a Lδe Lδ r ⎤
containing more than one state variable must be parameterized ⎢
s
⎥ ⎢ ⎥ (40)
a a22 a23 M bias ⎥ ⎢ 0 M δe 0 ⎥
1
A(x, σ1 ) = ⎢ 21 s
, B=⎢ ,
and apportioned among the corresponding elements of the ⎢ a31 a32 a33 1
Nbias ⎥ N Nδ e Nδ r ⎥

s
⎥ ⎢ δa ⎥
A(x, σ ) matrix (Cloutier & Stansbery, 2002b). For example, ⎣0 0 0 −λs ⎦
⎣⎢ 0 0 0 ⎦⎥
if x3 = x1 x2 , A(x) = [aij ] is parameterized with a31 = σ 1 x2 where the aij elements of the inner-loop SDC matrix
and a32 = (1 − σ 1 ) x1 . The free design parameter σ 1 can be A(x, σ1 ) are presented in Appendix A.1. State-dependent
selected to have any finite real value, but is generally chosen nonlinear formulation of the inner loop is then based on the
in the range [0,1]. It is also desirable to shift state-dependent
complete set of dynamic equations (28), (30) and (40).
factors which exclude the origin even though they are
Therefore, for control design, the inner loop is represented in
embedded in a term which goes to zero as the state goes to
SDC form by the augmented system
zero, as discussed in Cloutier & Stansbery (2002a,b) and
x IL = A IL (x IL )x IL + B IL u IL (41)
Çimen (2010). For example, consider x2 = x3 cos x1 .
Obviously, this term goes to zero as x3 goes to zero, but it is with the respective state and control input vectors
desirable to have a nonzero entry in the (2,1)-element of the x IL [xT δT δT ]T , u IL δ com .
A(x, σ ) matrix that reflects the fact that x2 depends on x1 . Using the relationship u = Δ logic δ from (28), together with
This is accomplished by shifting the term so that it goes (30) yields the SDC matrices for the inner loop as
through the origin. For the given example, adding and ⎡ ⎤
⎢ A(x, σ1 ) BΔ logic 04× 4 ⎥ (42)
subtracting 1 gives cos x1 = [cos x1 − 1] + 1 . The function A IL (x IL ) ⎢ 0 0 4× 4 I 4× 4 ⎥, B ⎡ 08× 4 ⎤
⎢ω 2I ⎥
cos x1 − 1 goes through the origin and can therefore be ⎢ ⎥ IL
4× 4
⎣ n 4× 4 ⎦
factored as cos x1 − 1 = [(cos x1 − 1) x1 ] x1 . Note that the
⎢ 0
⎣ 4× 4 −ωn 2 diag {
sat(δ i ,δ max )
δi } −2ζωn 2 diag { sat(δ i ,δ max )
δi } ⎥

expression in brackets is well-behaved at the origin, when for i = 1, … , 4 , with A(x, σ1 ) and B constructed as in (40).
x1 = 0 . Then, by writing x2 = [(cos x1 − 1) x1 ] x1 x3 + x3 allows
State-dependent nonlinear formulation of the inner-loop given
the system to be parameterized as a21 = σ 2 [(cos x1 − 1) x1 ] x3
by (41) and (42) incorporates hard bounds on the control
and a23 = (1 − σ 2 )(cos x1 − 1) + 1 , which yields the desired
surface deflection angles δ i and their rates δ i . This is
nonzero entry in a21 . Such an approach provides a complete
accomplished using the second-order nonlinear actuator
characterization of the possible factorizations of f (x) into
A (x, σ )x , where σ = [σ 1 , … , σ k ] . dynamics given by (30). This is different from the inner-loop

9597
18th IFAC World Congress (IFAC'11)
Milano (Italy) August 28 - September 2, 2011

design of the missile control problem presented by Cloutier & ⎡ 0 0 a13 ⎤ ⎡ b11 b12 ⎤
(44)
Stansbery (2001), where zero-order (that is, lag-free) actuator AOL (xOL ) = ⎢⎢ 0 0 a23 ⎥⎥ , BOL = ⎢⎢b21 b22 ⎥⎥ ,
⎢⎣ a31 0 0 ⎥⎦ ⎣⎢ 0 0 ⎦⎥
dynamics was assumed with hard bounds imposed only on the
effective control surface deflections angles ( δ a , δ e , δ r ) using with the resultant coefficients defined in Appendix A.3.
integral control. In reality, however, these constraints take The second alternative formulation follows Cloutier &
effect on individual control surfaces δ i ( i = 1, … , 4 ), which Stansbery (2002a), using angle of attack and bank angle
makes the approach proposed here more practical. Note also commands, while issuing a zero sideslip command for a BTT
that without actuator dynamics, imposing hard bounds on both autopilot. However, the controls α and μ appear nonlinearly
δ i and δ i would require the use of double integral control. in (36) as products of sines and cosines. Therefore, in order to
bring the dynamics into control-affine form (38), integral
5.2 State-Dependent Formulation of the Intermediate Loop control must be applied. Defining the state and control vectors
For state-dependent nonlinear formulation of the intermediate xOL [γ χ h α μ ]T and uOL [α μ ]T , bias terms
loop, let us define the state vector and the control vector as contained in (36) can be handled by multiplying and dividing
x ML [a X B aYB aZ B M α β μ ]T , u ML [ p q r ]T . these terms by the state h . The SDC parameterization can
then be constructed in a similar manner as before. Although
Since p , q , r are the controls in the intermediate loop, the this alternative formulation increases the order of the state
current values of the control surface deflections δ j and their vector by two, it can provide the desired set of controls for
rates δ j ( j = a, e, r ) are used in this loop. BTT missiles using flight path angle commands from the
Since thrust is not controllable, the controllability of the Mach guidance system (Cloutier & Stansbery, 2002a). The former
dynamics (25) is due to angle of attack and sideslip angle. formulation, however, has the advantage of providing
Consequently, Mach controllability is lost when α and β acceleration commands to the intermediate loop, which are
are both zero, as also noted by Cloutier & Stansbery (2001). directly available from sensor measurements, as opposed to
Therefore, similar to Cloutier & Stansbery (2001), a estimates of incidence angles required for the second
stabilizing term of −0.1M is added to the Mach dynamics so formulation. Therefore, the former formulation proposed here
as to eliminate loss of controllability in the state-dependent is used with the three-loop topology, whereas, for BTT
formulation of the intermediate-loop dynamics (Çimen, 2010). missiles, use of the inner two-loop structure is proposed.
Eqs. (25), (26), (27), (32) and (34) can now be represented in 6. CONTROLLER DESIGN
SDC form (39). Once again, following the systematic 6.1 Control Design Approaches
procedure previously described, any term containing more
than one state variable must first be parameterized and then For control design of the proposed three-loop structure
presented in Fig. 5, any extended linearization control method
apportioned among the corresponding elements of the A
(Çimen, 2010) such as pole placement can be applied.
matrix. Note that multiples of three individual states, say However, recent attention has been directed towards
x1 = x1 x2 x3 , can be parameterized as a11 = 13 (σ 1 + σ 2 ) x2 x3 , satisfying performance and stability robustness requirements
a12 = 13 (1 − σ 1 + σ 3 ) x1 x3 , a13 = 13 (2 − σ 2 − σ 3 ) x1 x2 . Therefore, a by designing autopilots using optimal control theory, because
vector σ 2 with 42 design parameters σ 2,1 , σ 2,2 , … , σ 2,42 ∈ these requirements for current and future missiles necessitate
is introduced, which is used in parameterizing nonlinear terms the use of optimally designed multivariable digital flight
in these equations. A hypersurface of SDC parameterizations control systems (Wise, 2007). LQR state feedback designs
for the intermediate-loop system is then constructed with generally give good performance characteristics and stability
⎡ a11 a12 a13 a14 0 0 0⎤ ⎡ Xp Xq Xr ⎤
⎢a ⎢ Y Yr ⎥⎥ (43) margins, with the availability of the states required for
⎢ 21 a22 a23 a24 0 0 0 ⎥⎥ ⎢ p Yq
⎢ a31 a32 a33 a34 0 0 0⎥ ⎢ Zp Zq Zr ⎥ implementation. Command tracking is achieved through this

A ML (x ML , σ 2 ) = ⎢ a41 a42 a43 −0.1 a45 a46
⎥ ⎢
0 ⎥ , B ML (x ML ) = ⎢ 0 0 0 ⎥,
⎥ framework by augmenting the system with an integrator, and
⎢ a51

0 a53 0 a55 0 0⎥

⎢ − cα t β

1 − sα t β ⎥

designing the controller for the augmented system. In many
⎢ a61 a62 a63 0 a65 a66 0⎥ ⎢ sα
⎢c c
0 − cα ⎥ practical designs not all the states are available for feedback
⎢a a72 a73 0 a75 a76 ⎥
a77 ⎦ 0 sα c β ⎥⎦
⎣ 71 ⎣ α β and, for these problems, H2 and H∞ frameworks can be
where the aij elements of the intermediate-loop SDC matrix applied. All these methods lead to the various State-
A ML (x ML , σ 2 ) are given in Appendix A.2. Dependent Riccati Equation (SDRE) control methodologies
proposed in the literature, such as the SDRE nonlinear
5.3 State-Dependent Formulation of the Outer Loop regulator, SDRE integral servomechanism, SDRE nonlinear
The outer loop is formulated using the flight path angle and H2 control, and SDRE nonlinear H∞ control, which are all
altitude dynamics derived in (36) and (37), respectively. viable approaches for controller design. These methods use
either state feedback or output feedback architectures, and are
However, there are two possible ways of formulating the outer
all the product of optimal control based designs.
loop in SDC form, which are now both discussed.
Using SDRE nonlinear regulation with integral control, the
For the first formulation, the state vector and the control SDRE controller can be implemented as an integral
vector for the outer loop are defined as servomechanism to provide a type 1 controller for all three
xOL [γ χ h]T , u OL [aYB aZ B ]T . loops in order to perform perfect tracking of the loop
The bias terms contained in (36) are handled by multiplying commands (Çimen, 2010). This is accomplished by first
decomposing the state vector x (which must be available for
and dividing these terms by the state h , which never goes to
feedback) as xT = [x RT x N T ] , where it is desired for the
zero. Constructing the required SDC parameterization is then
vector components of x R (the tracked states) to track a
very straightforward, giving reference, input or command signal r ∈ p , and x N consists

9598
18th IFAC World Congress (IFAC'11)
Milano (Italy) August 28 - September 2, 2011

of the nontracked states. The state vector x is then augmented aerodynamic normal force is in the desired direction, and the
with x I , the integral states of x R . The augmented system magnitude of the force is controlled by adjusting the angle of
becomes attack in that plane.
x = A (x, σ )x + B(x)u, Recall that maximizing overall missile performance requires
⎡0 I 0 ⎤ ⎡ 0 ⎤ choosing the appropriate autopilot command structure for
x [x I T
xR T
xN ] , A ( x, σ ) = ⎢
T T
⎥ , B ( x) = ⎢ B ( x) ⎥
⎣ 0 A(x, σ ) ⎦ ⎣ ⎦ each mission phase. The genericity of the design presented in
with A(x, σ ) and B(x) derived in Section 5 for each of the Sections 4 and 5 thus becomes apparent, where the reference
three loops. The problem is then formulated as a nonlinear command can be selected as any of the state variables,
optimal control problem, with the minimization of depending on the specific control mode, intercept scenario, or
∫ {x Q ( x ) x + u }

J= 1
2
T T
R (x)u dt. mission phase. For example, for STT missiles, the tracking
0
command is usually given by r = [aYcom aZcom μ com ]T , so that
Then, by mimicking the LQR formulation, the full-state x R is defined as x R = [aYB aZ B μ ]T , with μ com = 0 such
B B

feedback, robust SDRE integral servo controller becomes that the bank angle is regulated to zero. For BTT missiles, on
⎡ x I − ∫ r dt ⎤
u ( x ) = − R −1 ( x ) B T ( x ) P ( x ) ⎢ x R − r ⎥ , the other hand, the autopilot can easily be transformed to
⎢ xN ⎥ follow roll angle and angle-of-attack commands generated by
⎣ ⎦
where P (x) is the unique, symmetric, positive-definite the guidance system (or the outer loop, if desired), so that
(pointwise stabilizing) solution of the SDRE r = [α com β com μcom ]T and thus x R = [α β μ ]T with
P ( x ) A ( x , σ ) + A T ( x , σ ) P ( x ) − P ( x ) B ( x ) R −1 ( x ) B T ( x ) P ( x ) + Q ( x ) = 0. β com = 0 , since the sideslip angle must be kept as small as
possible during flight. In case of altitude-hold autopilots, the
In the absence of the integral states x I , the above formulation
altitude in the outer loop may be selected as the commanded
reduces to the SDRE servomechanism problem. input, so that r = h com and xR = h . The required autopilot
command structure then becomes a consequence of the chosen
6.2 Selection of Design Parameters
reference commands. If body rates are the guidance
The respective vector design parameters σ1 and σ 2 defining commands, then only the single inner-loop layer in Fig. 5 is
the hypersurface of SDC parameterizations given in (40) and required for autopilot design. Alternatively, for acceleration
(43) (see Appendix A) must now be selected for controller or incidence angle commands, the required autopilot topology
design. A systematic and an effective procedure has been is given by the inner two-loop structure of Fig. 5. On the other
described in Çimen (2010) for selecting the “best” SDC hand, if the guidance algorithm supplies flight path angle or
parameterization, which is to attempt to maximize the altitude commands, the complete three-loop autopilot design
pointwise controllable spaces of the possible factorizations by topology of Fig. 5 is then required. Additionally, several
computing the state-dependent controllability matrix possible combinations can be used for each mission phase by
M C ( x, σ i ) = [ B ( x ) A ( x, σ i ) B ( x ) A n −1 (x, σ i )B(x)], i = 1, 2 switching between the required command structures within
and evaluate it for different values of σ1 and σ 2 . This is the proposed topology during different phases of flight.
clearly the logical choice, since pointwise control effort can
be directly linked to these issues (for a detailed discussion 6.4 Performance and Robustness Results
with examples, readers may refer to Çimen, 2010). The genericity, performance and robustness of the autopilot
In order for the SDRE to have a solution, the pointwise design approach developed in the paper has been tested and
detectability condition must also be satisfied (Çimen, 2010). evaluated on several alternative missile configurations, using
The state-dependent weighting matrix Q(x) must be selected different intercept scenarios, and for entire engagements, from
and tuned for the entire range of flight conditions, in order to launch to intercept, using detailed 6-DOF simulations and
maintain good autopilot response over the entire flight Monte Carlo runs. In particular, the developed approach has
envelope (Cloutier & Stansbery, 2001). The state weightings been used in autopilot design for several hypothetical missile
on the tracked outputs, therefore, must be varied at low, configurations, including a highly unstable missile airframe
medium and high speeds and altitudes until similar with canard controls, stable and unstable nonminimum-phase
performance is achieved for all cases. The state weightings tail-controlled configurations, and a configuration which has
can then be curve fit to a quadratic function of dynamic combined characteristics, with a separable booster mechanism
pressure q , which is a function of both altitude and airspeed, that switches between thrust vectoring during launch and
and hence Mach number (one of the states in the outer loop). canard control after separation. In the latter case, different
Alternatively, they can be calculated online by interpolation. autopilots for vertical launch, midcourse and terminal phases
have been successfully implemented for entire engagements,
6.3 Design Genericity based on different sets of autopilot command structures in Fig.
5. The nonlinear autopilot has been implemented into the
Skid-to-turn (STT) and bank-to-turn (BTT) are the two basic simulation and integrated with a high-order control actuation
modes of controlling a missile. In the STT mode, the roll system, as opposed to the second-order dynamics assumed in
angle is either held constant or uncontrolled. This has the the autopilot model. Guidance and navigation algorithms were
major advantage of decoupling roll dynamics from the pitch supplied using a strapdown IMU sensor, taking into account
and yaw dynamics, and delivering faster response than the sensor location. Simulations were then performed with the
BTT mode. A BTT missile, on the other hand, allows only nonlinear autopilot solved online at a rate of 1 kHz. Due to
positive angles of attack while maintaining small sideslip space limitations, however, the various sets of simulation
angles to prevent missile maneuvers from shading the inlet of results could not be included in the paper, and will be
air-breathing engines in order to increase fuel efficiency and presented during Congress presentation. At this point,
thereby maximize range. The desired orientation is achieved however, it is worth mentioning that the inner-, intermediate-
by rolling (banking) the missile so that the plane of maximum and outer-loop SDRE controllers all provide excellent

9599
18th IFAC World Congress (IFAC'11)
Milano (Italy) August 28 - September 2, 2011

tracking performance over the full range of flight conditions REFERENCES


and throughout the entire operating envelope, for various J. H. Blakelock (1991). Automatic Control of Aircraft and Missiles, 2nd Ed.,
missile configurations and intercept scenarios. This should not Wiley.
come as a surprise, since outstanding performance of SDRE- T. Çimen (2010). Systematic and effective design of nonlinear feedback
controllers via the state-dependent Riccati equation (SDRE) method.
based designs for nonlinear missile autopilots have already Annual Reviews in Control, 34, 32–51.
been demonstrated in several works in the literature (see, for J. R. Cloutier, and D. T. Stansbery (2001). Nonlinear, hybrid bank-to-
turn/skid-to-turn autopilot design. In: Proc. of the AIAA Guidance,
example, Mracek & Cloutier, 1996, 1997; Cloutier & Navigation, and Control Conference, Montreal, Canada.
Stansbery, 2001, 2002a; and Vaddi, Menon & Ohlmeyer, J. R. Cloutier, and D. T. Stansbery (2002a). Dynamic conversion of flight
path angle commands to body attitude commands. In: Proc. of the
2009). Simulations have shown that at different flight American Control Conference, Anchorage, AK, USA, 221–225.
conditions within the flight envelope the time-scaled J. R. Cloutier, and D. T. Stansbery (2002b). The capabilities and art of state-
dependent Riccati equation-based design. In: Proc. of the American
nonlinear autopilot structure presented in Fig. 5 consistently Control Conference, Anchorage, AK, USA, 86–91.
delivers uniform performance when designed with SDRE B. Etkin (2005). Dynamics of Atmospheric Flight, Dover Books on
control. Thus, the SDRE autopilot provides the same level of Engineering.
J. L. Meriam, and L. G. Kraige (1986). Engineering Mechanics, Volume 2,
performance at all Mach numbers and altitudes. Additionally, Dynamics, Second Edition, John Wiley & Sons, Inc., New York, NY.
since no gain scheduling is required to implement the C. P. Mracek, and J. R. Cloutier (1996). Missile longitudinal autopilot design
using the state-dependent Riccati equation method. In: Proc. of the
autopilot, tremendous savings in the design effort can be International Conference on Nonlinear Problems in Aviation and
realized. Moreover, robust performance of the nonlinear Aerospace, Daytona Beach, FL, USA, 387–396.
C. P. Mracek, and J. R. Cloutier (1997). Full envelope missile longitudinal
autopilot design approach using the SDRE method has been autopilot design using the state-dependent Riccati equation method. In:
evaluated and validated in the presence of high uncertainties Proc. of the AIAA Guidance, Navigation, and Control Conference, New
Orleans, LA, USA, 1697–1705.
in the aerodynamic coefficients, measurement inaccuracies, J. N. Nielsen (1988). Missile Aerodynamics, AIAA.
noise and delays, INS errors, as well as continuous wind and B. L. Stevens, and F. L. Lewis (2003). Aircraft Control and Simulation, 2nd
turbulent atmospheric effects, achieving outstanding Ed., John Wiley & Sons.
S. S. Vaddi, P. K. Menon, and E. J. Ohlmeyer (2009). Numerical state-
performance in all these cases. dependent Riccati equation approach for missile integrated guidance
control. AIAA Journal of Guidance, Control, and Dynamics, 32, 699–703.
7. CONCLUSIONS K. A. Wise (2007). A trade study on missile autopilot design using optimal
control theory. In: Proc. of the AIAA Guidance, Navigation and Control
In this study, a generic approach to missile autopilot design Conference and Exhibit, Hilton Head, South Carolina, USA.
K. A. Wise, and D. J. Broy (1998). Agile missile dynamics and control. AIAA
has been developed using state-dependent nonlinear control. Journal of Guidance, Control, and Dynamics, 21, 441–449.
The proposed topology has a three-loop structure. The outer P. H. Zipfel (2007). Modeling and Simulation of Aerospace Vehicle
Dynamics, AIAA.
loop uses either flight path angle commands or altitude
commands from the guidance law to generate roll angle Appendix A. SDC MATRIX COEFFICIENTS
command together with either body acceleration commands or
A.1 Inner-Loop SDC Matrix Coefficients
incidence angle commands, depending on whether the missile
a11 = Lp + σ 11 Lpq q a12 = M p + M p 2 p + σ 13 M pr r a13 = N p + σ 14 N pq q
flies STT or BTT. The intermediate loop tracks the outer-loop
a21 = (1 − σ 11 ) Lpq p + σ 12 Lqr r a22 = M q a23 = (1 − σ 14 ) N pq p + σ 15 N qr r
commands by generating commanded body (roll, pitch and
a31 = Lr + (1 − σ 12 ) Lqr q a32 = M r 2 r + (1 − σ 13 ) M pr p a33 = N r + (1 − σ 15 ) N qr q
yaw) rates. The inner loop then tracks the intermediate loop
body rate commands using the control surface deflections. In A.2 Intermediate-Loop SDC Matrix Coefficients
this topology, any extended linearization control method, such a11 = X aX σ 2,1 M a12 = X aY σ 2,2 M a13 = X aZ σ 2,3 M
as various SDRE methods, can be applied in the design of all a21 = Ya X σ 2,4 M a22 = YaY σ 2,5 M a23 = YaZ σ 2,6 M
three loops. Full nonlinear and coupled 6-DOF equations of a31 = Z aX σ 2,7 M a32 = Z aY σ 2,8 M a33 = Z aZ σ 2,9 M
motion describing missile dynamics are derived and used a14 = X aX (1 − σ 2,1 )a X B + X aY (1 − σ 2,2 )aYB + X aZ (1 − σ 2,3 )aZ B + X M 2 M
directly in all three loops, where state-dependent nonlinear a24 = YaX (1 − σ 2,4 )a X B + YaY (1 − σ 2,5 )aYB + YaZ (1 − σ 2,6 )aZ B + YM 2 M

control is applied in each loop without making any a34 = Z aX (1 − σ 2,7 )a X B + Z aY (1 − σ 2,8 )aYB + Z aZ (1 − σ 2,9 )aZ B + Z M 2 M

simplifications to the equations. A general form of the a41 = 1a [ 13


cα −1 cβ −1
α β (σ 2,10 + σ 2,11 )αβ + (cα − 1)σ 2,13 + (cβ − 1)σ 2,14 + 1]

hypersurface of SDC parameterizations is constructed for the a42 = σ 2,15 s β


1
a

a43 = 1a [σ 2,16 sα + 13 (cβ − 1)(α 2,17 + α 2,18 )sα ]


inner, intermediate and outer loops to carry out the control cα −1 cα −1
(cβ − 1)(1 − σ 2,10 + σ 2,12 )aX B + (1 − σ 2,13 )aX B + (1 − σ 2,16 )aZ B + 13 (cβ − 1)(2 − α 2,18 − α 2,19 )aZB
sα sα
a45 = 1a [ 13
design. This topology also preserves the inherent time-scale α
cβ −1
α
cβ −1
α α ]
sβ 1 cβ −1
separation between translational and rotational dynamics of a46 = [ (cα − 1)
1 1
a 3 β
(2 − σ 2,11 − σ 2,12 )aX B + β
(1 − σ 2,14 )aX B + (1 − σ 2,15 )aYB β
+ 3 β
(1 − α 2,17 + α 2,19 )aZB sα ]
cα −1
the short-period mode, and produces an effective approach to a51 = − Ma1cβ σ 2,21 sα [1 + (cα − 1)σ 2,20 ] a55 = (1 − σ 2,20 )aZ B − (1 − σ 2,21 )aX B

a53 = 1
Ma cβ
1
Ma cβ
[ α α ]

missile autopilot design. Detailed 6-DOF simulations have a61 = − 1 1


Ma 3
[ (cα − 1)(σ 2,22 + σ 2,23 )sβ + σ 2,25 s β ]
a62 = [(cβ − 1)σ 2,26 + 1]
also indicated that the proposed approach significantly
1
Ma

a63 = − 13 Ma
1
(σ 2,27 + σ 2,28 )sα sβ
improves gain-scheduled linear autopilots. The generic cα −1
(1 − σ 2,22 + σ 2,24 )aX B + (1 − σ 2,27 + σ 2,29 )aZB

a65 = − 13 Ma
1
sβ [ ]
approach developed in this study for autopilot design can be α
sβ sβ
α
cβ −1 sβ
a66 = − Ma [ 3 (cα − 1)(2 − σ 2,23 − σ 2,24 )aX B + (1 − σ 2,25 )aX B − (1 − σ 2,26 )aYB + 13 (2 − σ 2,28 − σ 2,29 )aZB sα
easily extended to even more general cases, such as advanced ]
1 1
β β β β

a71 = 1
[ (α 2,30 + α 2,31 ) sα t β + (α 2,33 + α 2,34 ) sα t γ s μ − σ 2,36 cα s β t γ c μ ]
1 1
missile control problems with thrust vectoring and reaction Ma 3 3

a72 = 1
cβ tγ cμ , a73 = 1
[ − 13 (α 2,37 + α 2,38 ) sα s β t γ c μ + σ 2,41 cα t β + σ 2,42 cα t γ s μ ]
jets, and even other flight control design problems using the Ma Ma

[ 13 (1 − α 2,30 + α 2,32 ) a X B t β + 13 (1 − α 2,33 + α 2,35 ) a X B t γ s μ − 13 (1 − α 2,37 + α 2,39 ) aZ B s β t γ c μ ] αα


s
a75 = 1
general form of the 6-DOF rigid-body equations of motion in Ma

a76 = 1
[ (2 − α 2,31 − α 2,32 ) a X B s
1 1
− (1 − σ 2,36 ) a X B cα t γ c μ − (2 − α 2,38 − α 2,39 ) a Z B sα t γ c μ
1

the design of state-dependent nonlinear flight control systems. Ma 3


α cβ 3

+ (1 − σ 2,41 ) a Z B cα 1

] β

ACKNOWLEDGEMENTS a77 = 1
Ma
[ 13 (2 − α 2,34 − α 2,35 ) a X B sα + (1 − σ 2,42 ) a Z B cα ] t γ

μ

The author would like to thank Prof. M. Kemal Özgören


A.3 Outer-Loop SDC Matrix Coefficients
(Middle East Technical University, Turkey) and
Mr. Bahadır Çor (ROKETSAN Missiles Industries Inc., a13 1
Mah
[(sα c μ + cα s β s μ )a X B + g1 ] b11 − Ma
1
cβ sμ b12 − Ma
1
(cα c μ − sα sβ s μ )
[(sα s μ − cα s β cμ )a X B + g 2 ] b21 − Ma1cγ (cα s μ + sα s β c μ )
Turkey) for providing constructive comments and valuable a23 cβ cμ b22
1 1
Mah cγ Ma cγ

a31 Ma sγ γ
ideas, which helped significantly improve this paper.

9600

You might also like