Download as pdf or txt
Download as pdf or txt
You are on page 1of 101

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/382125192

Advanced Strategies for Guidance & Control of Surface-Air Missiles

Book · July 2024

CITATIONS READS

0 50

1 author:

Belkacem Bekhiti
Saad Dahlab University
60 PUBLICATIONS 145 CITATIONS

SEE PROFILE

All content following this page was uploaded by Belkacem Bekhiti on 11 July 2024.

The user has requested enhancement of the downloaded file.


Advanced Strategies for
Guidance & Control of
Surface-Air Missiles

Dr. BEKHITI BELKACEM


Institute of Aeronautics and Space Studies
University of Saad Dahleb-Blida
Academic year: 2024-2025
Summaries

A
bstract: Over the last three decades, there have been many
studies in the area of missile guidance and control. The result has
been a great deal of progress and several approaches to the problem
have emerged. The basic problem is to intercept a target with great
accuracy in an environment that is uncertain and noisy. One of the
earliest forms of missile guidance is that of pure pursuit, command to line of sight
and the proportional navigation, this involves establishing a line of sight (LOS)
between the tracking sensor and the target. This work investigates the guidance
and control design problem for a generic surface to air missile intercepting a given
target using different optimized guidance laws which are the: optimized pure
pursuit, optimized commend to line of sight, intelligent proportional navigation,
and optimized proportional derivative based guidance. The performances of these
guidance laws are tested against a given target in terms of the achieved miss-
distance and the time of closest approach. Furthermore, qualitative comparative
study between the aforementioned guidance laws is presented.

-------------------------- ooo --------------------------

‫الملخص‬
‫ اكنت هناك العديد من ادلراسات يف جمال التوجيه والس يطرة الصاروخية‬،‫عىل مدى العقود الثالثة املاضية‬
‫ وقد أسفرت هذه ادلراسات عن‬.‫حبيث أجريت العديد من الحباث يف جمال توجيه الصوارخي والتحمك فهيا‬
‫ تمتثل املشلكة الساس ية يف اعرتاض الهدف بدقة‬.‫تقدم كبري وظهور عدة طرق للتعامل مع هذه املشلكة‬
‫ التوجيه عن‬:‫ انه من بني أقدم أشاكل التوجيه الصارويخ هو‬.‫عالية ويف بيئة غري مؤكدة ومليئة ابلضوضاء‬
،‫ و التوجيه عن طريق التحمك يف خط الرؤية‬،‫طريق املطاردة اخلالصة وتسمى أيضا التتبع البرصي املبارش‬
.‫ وهذا يتضمن انشاء خط رؤية بني مستشعر التتبع والهدف‬،‫و التوجيه عن طريق املالحة التناسبية‬
‫ جو وليك يعرتض هدفًا معينًا‬-‫يبحث هذا العمل يف مشلكة تصممي التوجيه والتحمك العام لصاروخ أرض‬
‫ املطاردة اخلالصة احملس نة أو ما يسمى أيضا بــ التتبع البرصي‬:‫وحمس نة ويه‬
ّ ‫ابس تخدام قوانني توجيه خمتلفة‬
‫ والتوجيه املش تق التناس يب‬،‫ واملالحة التناسبية احملسن‬،‫ والتوجيه اىل خط الرؤية احملسن‬،‫املبارش احملسن‬
‫ يمت اختبار أداء قوانني التوجيه هذه مقابل هدف معني من ييث مسافة اخلط احملققة ووقت‬.‫احملسن‬
.‫ مت عرض دراسة نوعية مقارنة بني القوانني الرشادية املذكورة أعاله‬،‫ عالوة عىل ذكل‬.‫القرتاب القرب‬
Notations and Acronyms

𝐏 Position vector from the center of mass. 𝑻1 (𝜙) Rotation about the xb-axis through 𝜙.
𝐯(𝑡) Absolute linear velocity of rigid body. 𝑻2 (𝜃) Rotation about the yb-axis through θ.
𝐯̇ (𝑡) Absolute acceleration of center of mass. 𝑻3 (𝜓) Rotation about zb-axis through ψ.
∆𝐯 Change in missile velocity during ∆𝑡. 𝑻(𝜙, 𝜃, 𝜓) Rotation matrix about the 3 axes.
𝑳𝑖 Initial total system momentum. 𝐢𝑏 , 𝐣𝑏 , 𝐤 𝑏 Unit vectors of body fixed frame.
𝑳f Final total system momentum. 𝑥, 𝑦, 𝑧 Axes of coordinate system.
𝛚(𝑡) Angular velocity of the rigid body. 𝑥𝐵 , 𝑦𝐵 , 𝑧𝐵 Coordinates of the body frame.
𝛚̇(𝑡) Angular acceleration of the rigid body. 𝑥𝐸 , 𝑦𝐸 , 𝑧𝐸 Coordinates of the Earth frame.
𝐯𝑒 (𝑡) Absolute velocity of exhaust gases. 𝑥𝑊 , 𝑦𝑊 , 𝑧𝑊 Coordinates of the wind frame.
𝐯𝑟𝑒 (𝑡) Relative velocity of the exhaust gases. 𝐶𝐹 general aerodynamic force coefficient.
𝐮𝑣𝑒 Unit vector in direction of 𝐯𝑟𝑒 . 𝑆 aerodynamic reference area.
𝐋(𝑡) Linear momentum of the total system. 𝜌 atmospheric density.
𝐅(𝑡) General force (aerodynamic [N]). 𝑀𝑁 Mach number 𝑀𝑁 = 𝑉𝑚 /𝑉𝑠 .
𝐅𝑒𝑥𝑡 (𝑡) Sum of forces acting on the missile. 𝑃𝑎 ambient atmospheric pressure.
𝐅𝐴 (𝑡) Resultant aerodynamic force vector. 𝑄 dynamic pressure parameter.
𝐅g (𝑡) Gravitational force vector. 𝑉𝑠 speed of sound at altitude ℎ.
𝐅𝑃 (𝑡)Total instantaneous thrust force vector 𝑅 gas constant (287.05).
𝐇(𝑡) The total angular momentum. 𝑇 temperature at altitude.
𝐌(𝑡) Total moment acting on particle/body. 𝛾 ratio of specific heat (1.4).
𝐌𝐴 (𝑡) Aerodynamic moment of the system. 𝑅𝑒 Reynolds Number.
𝐌𝑃 (𝑡) Thrust moment of the system. 𝑑 aerodynamic reference length of body.
[𝑰] Inertia matrix of a body. 𝜇 atmospheric dynamic viscosity.
𝐹𝐴𝑥 , 𝐹𝐴𝑦 , 𝐹𝐴𝑧 Components of 𝐅𝐴 in body frame 𝐶𝐷 aerodynamic drag coefficient.
𝐹𝑃𝑥 , 𝐹𝑃𝑦 , 𝐹𝑃𝑧 Components of 𝐅𝑃 in body frame 𝐶𝐷0 zero-lift drag coefficient.
𝐹g𝑥 , 𝐹g𝑦 , 𝐹g𝑧 Components of 𝐅g in BF 𝐶𝐿 aerodynamic lift coefficient.
𝑀𝑥 , 𝑀𝑦 , 𝑀𝑧 Components of 𝐌. 𝐶𝐿𝛼 slope of curve formed by 𝐶𝐿 versus 𝛼 .
𝐶𝑙 aerodynamic roll moment coefficient.
𝐿𝐴 , 𝑀𝐴 , 𝑁𝐴 Components of 𝐌𝐴 .
𝐶𝑚 aerodynamic pitch moment coefficient.
𝐿𝑃 , 𝑀𝑃 , 𝑁𝑃 Components of 𝐌𝑃 .
𝐶𝑛 aerodynamic yaw moment coefficient.
𝐻𝑥 , 𝐻𝑦 , 𝐻𝑧 Components of the 𝐇.
𝐶𝑛𝛽 slope of curve formed by 𝐶𝑛 versus 𝛽.
𝐼𝑥𝑥 , 𝐼𝑦𝑦 , 𝐼𝑧𝑧 Components of diagonal [𝑰].
𝐶𝑛𝛿 slope of curve formed by 𝐶𝑛 versus 𝛿𝑦 .
𝐼𝑥𝑦 , 𝐼𝑥𝑧 , 𝐼𝑦𝑧 Products of inertia.
𝐶𝑚𝛼 slope of curve formed by 𝐶𝑚 versus 𝛼.
𝑢, 𝑣, 𝑤 Components of 𝐯 in the body frame.
𝐶𝑚𝛿 slope of curve formed by 𝐶𝑚 versus 𝛿𝑝 .
𝑢̇ , 𝑣̇ , 𝑤̇ Components of 𝐯̇ in the body frame.
𝐶𝑛𝑎 slope of curve formed by 𝐶𝑛 versus 𝛿𝑝 .
𝑝, 𝑞 , 𝑟 Components of 𝛚 in body frame.
𝑝̇ , 𝑞̇ , 𝑟̇ Components of 𝛚̇ in body frame. 𝛿𝑝 deflection angle in the pitch direction.
𝜙, 𝜃, 𝜓 Euler angles in roll, pitch and yaw. 𝛿𝑟 deflection angle in the roll direction.
𝜙̇, 𝜃̇, 𝜓̇ Rate of change of Euler angles. 𝛿𝑦 deflection angle in the yaw direction.
𝐹𝑆 Magnitude of aerodynamic side force 𝑭𝑆 . 𝐶𝑚ref pitching moment coefficient.
𝐹𝐷 Magnitude of aerodynamic drag force 𝑭𝐷 . 𝐶𝑛ref yawing moment coefficient.
𝐹𝐿 Magnitude of aerodynamic lift force 𝑭𝐿 . 𝐶𝑁y coefficient of normal force on yb axis.
𝐴 Magnitude of aerodynamic axial force 𝑨. 𝐶𝑁z coefficient of normal force on zb axis.
𝑁 Magnitude of aerodynamic normal force 𝑵
𝑷𝑀 position vector of the missile, [m]. 𝐶𝑛𝑟 yaw damping derivative relative to 𝑟.
𝑷 𝑇 position vector of the target, [m]. 𝐶𝑛 yaw damping derivative relative to 𝛽̇ .
𝛽̇
𝐫 range between missile and target, 𝐶𝑚𝑞 pitch damping derivative relative to 𝑞 .
𝐯𝑀 interceptor missile velocity,
𝐶𝑚𝛼̇ pitch damping derivative relative to 𝛼̇ .
𝐯𝑇 velocity of the target,
𝐶𝑙𝑝 roll damping derivative, [rad−1 ]
𝛼 Angle of attack in pitch plane.
𝛽 Angle of sideslip. 𝐶𝑙𝛿 slope of curve formed by 𝐶𝑙 versus 𝛿𝑟 .
𝛼𝑡 Total angle of attack. 𝑃 pressure at altitude ℎ.
𝐴𝑒 Rocket nozzle exit area. 𝑃1 pressure at given altitude ℎ1 .
𝑚(𝑡) Instantaneous mass of missile. 𝛿𝑃 autopilot pitch fin command,
𝑚̇(𝑡) Rate of change of missile mass. 𝛿𝑅 autopilot roll fin commands.
𝑚̇𝑒 (𝑡) Mass rate of flow of exhaust gas. 𝛿𝑌 autopilot yaw fin command,
𝑝𝑎 Ambient atmospheric pressure. 𝛿𝑖 deflection angle of 𝑖𝑡ℎ control surface.
𝑝𝑒 Average pressure at nozzle exit area. 𝑠 the Laplace variable.
𝐠 Acceleration vector-due-to-gravity. 𝛿(𝑠) achieved control-surface deflection, rad
g Magnitude of acceleration 𝐠. 𝛿𝑐 (𝑠) commanded control-surface deflection
𝐺 Gravity constant 6.673 × 10−11 m3 (kg/s2 ). 𝐺(𝑠) Control system transfer function.
ℎ Altitude above sea level. 𝐾 servo system gain, [s−1 ]
𝜆 line-of-sight (LOS) angle [rad], 𝑥cm : distance from nose to center mass [m]
𝑑𝜆/𝑑𝑡 the LOS rate [rad/sec]. 𝑥ref : distance from nose to reference point.
𝛾𝑚 missile flight path (or heading) angle. 𝐼𝑠𝑝 specific impulse of propellant [N.s/kg]
𝛾𝑡 target flight path angle. Fpref (𝑡) reference thrust force, [N]
PN Proportional navigation. 𝐮gl guideline unit vector.
MIMO Multiple input multiple output. 𝛚gl angular velocity of the guideline [rad/s].
Mag[ ] magnitude of the argument vector. 𝒂𝑐 commanded normal acceleration [m/s 2 ]
𝑁 navigation constant, positive real number. 𝐯𝑐 the closing velocity [m/s],
• Diff-classes of guided missiles
General Introduction • Types of guidance systems
to Guided Missiles • Computer simulation techniques

A guided missile is an unmanned vehicle that travels above the earth's surface; it
carries an explosive war head or other useful payload; and it contains within itself
some means for controlling its own trajectory or flight path. A glide bomb is
propelled only by gravity. But it contains a device for controlling its flight path, and
is therefore a guided missile. A missile is any object that can be projected or
thrown at a target. This definition includes stones and arrows as well as gun
projectiles, bombs, torpedoes, and rockets. But in current military usage, the word
missile is gradually becoming synonymous with guided missile. It will be so used in
this text; we will use the terms missile and guided missile interchangeably.
Security requirements prevent any detailed description of specific missiles in an
unclassified text. This text will therefore contain no information about specific
missiles; they will be described in some detail in a supplementary volume.

The reader will find some overlap and repetition in this text; this is intentional. The
subject is complex; it deals with many different phases of science and technology.
The beginning student of guided missiles faces a paradox. We might say that you
can't thoroughly understand any part of a guided missile unless you understand
all the other parts first. We will deal with this problem by first discussing the
guided missile as a whole, with a brief consideration of its dynamics, propulsion,
aerodynamics, control, guidance, and launching systems.

Modern military aircraft can fly so high and so fast that conventional antiaircraft
guns are ineffectual against them. As you know, a gun is not aimed directly at a
moving target; it must be so aimed that both the projectile and the target will
reach a predicted point at the same time. During the flight time of the projectile, a
high-speed high-altitude aircraft will travel several miles. Any slight change of
course during that time will take it beyond the lethal range of the projectile burst.

The surface-to-air guided missile is a more effective means of defense against


enemy aircraft. The missile can intercept attacking aircraft at greater heights, and
greater ranges, than any projectile. And the aircraft is unlikely to escape a missile
by taking evasive action. The missile is faster and more maneuverable. If the
attacking aircraft changes its course, the missile guidance system will change the
course of the missile accordingly, up to the instant of interception.

Guided missiles are becoming increasingly important in aircraft armament. When


two jet aircraft are approaching each other head-on, the range closes at a speed
between half a mile and one mile per second. Under these conditions it is difficult
even to see an enemy aircraft, and hitting it with conventional aircraft weapons
would be largely a matter of luck. But the air-to-air missile can "lock on" the
hostile aircraft while it is still miles away, and it can pursue and hit the target in
spite of its evasive maneuvers.
Guided missiles are classified in a number of different ways; perhaps most often by
function, such as air-to-air, surface-to-air, or air-to-surface. A non-ballistic missile
is propelled during all or the major part of its flight time; the propulsion system of
a ballistic missile operates for a relatively short time at the beginning of flight;
thereafter, the missile follows a free ballistic trajectory like a bullet (except that
this trajectory may be subject to correction, if necessary, by the guidance system).
Missiles may be further classified by type of propulsion system, such as turbo-jet,
ramjet, or rocket; or by type of guidance, such as command, beam-riding, or
homing. The missile guidance system keeps the missile on the course that will
cause it to intercept the target. It does this in spite of initial launching errors, in
spite of wind or other forces acting on the missile, and in spite of any evasive
actions that the target may take. The guidance system may be provided with
certain information about the target before launching. During flight it may receive
additional information, either by radio from the launching site or other control
point, or from the target itself. On the basis of this information, the guidance
system will calculate the course required to intercept the target, and it will order
the missile control system to bring the missile onto that course.

From the paragraph above, it might be inferred that the guidance system is an
intelligent mechanism that can think. This, of course, is untrue. The missile
guidance system is based on a relatively simple electronic computer. A computer
can take no action that isn't built into it by its designer (except, of course, the
erratic action that might result from a bad connection or a faulty component).
There are some types of missiles which are sensitive to infrared (heat) radiation,
and will steer itself toward any strong source of infrared. But an infrared is not the
only basis for homing guidance. A missile can also be designed to home on light,
radio, or radar energy given off by, or reflected from, the target. (It could also, like
a homing torpedo, be designed to home on a source of sound waves; but because a
guided missile travels at more than sound's speed, such a system would not be
practical). Because some source of missile's information is the energy given off by
the target itself, then its guidance is an example of passive homing. Other missiles
carry a radar transmitter, "illuminate" the target with a radar beam, and home on
the radar energy reflected from the target. This is an active homing guidance
system. A semi-active system is also possible; the target is illuminated by a radar
beam from the launching site or other control point, and the missile homes on
energy reflected from the target.

Although missiles are popularly known by their names, such as Sidewinder or


Terrier, every missile is assigned a designation consisting of letters and numerals.
The first three letters indicate the intended use of the missile:

AAM-air-to-air missile SAM-surface-to-air


ASM-air-to-surface SSM-surface-to-surface
AUM-air-to-underwater UAM-underwater-to-air
SUM-surface-to-underwater USM-underwater-to-surface
A guided missile, by definition, flies above the surface of the earth. Aerodynamic
long-range missiles, as well as all missiles of short and medium range, are subject
throughout their flight to the forces imposed by the earth's atmosphere. Ballistic
missiles, though they follow a trajectory that takes them into space, must climb
through the atmosphere after launching, and must descend through it before
striking the target. All missiles are subject to gravitational and inertial forces.

This first chapter will briefly discuss the principal forces that act on a guided
missile during its flight. It will show how the missile trajectory may be controlled
by designing the guidance law to affect the control surfaces. An understanding of
missile aerodynamics requires a familiarity with several of the basic laws of
physics. In such chapter you will find a detailed study of the rigid-body motion,
and the mathematical analysis of the various forces.

The resultant force on a wing can be resolved into forces perpendicular and
parallel to the relative wind; these components are lift and drag. The lift force
depends on the contour of the wing, the angle of attack, air density, area of the
wing, and the square of the air speed. If a missile is to continue in level flight, its
total lift must equal its weight. As the angle of attack increases, the lift increases
until it reaches a maximum value. At the angle of maximum lift, the air no longer
flows evenly over the wing, but tends to break away from it. This breaking away
(the burble point) occurs at the stalling angle. If the angle of attack is increased
further, both lifting force and airspeed decrease rapidly.

Drag is the resistance of air to motion through it. The drag component of the
resultant force on a wing is the component parallel to the direction of motion. This
force resists the forward motion of the missile. If the missile is to fly, drag must be
overcome by thrust-the force tending to push the missile forward. Drag depends on
the missile area, the air density, and the square of the velocity. Air resists the
motion of all parts of the missile, including the wings, fuselage, tail airfoils, and
other surfaces. The resistance to those parts that contribute lift to the missile is
called induced drag. The resistance to all parts that do not contribute lift is
parasitic drag. Because drag is proportional to the square of the speed, drag
increases very rapidly. The force of thrust is thus opposed by a steadily increasing
force of drag. The missile will continue to increase in speed, but its acceleration
(rate of increase of speed) will steadily decline. This decline will continue until
thrust and drag are exactly in balance; the missile will then fly at a uniform speed
as long as its thrust remains constant. If the propulsive thrust is decreased for any
reason (such as a command from the guidance system, or incipient fuel
exhaustion) the force of drag will exceed the thrust. The missile will slow down
until the two are again in balance. When the missile fuel is exhausted, or the
propulsion system is shut down by the guidance system, there is no more thrust.
The force of drag will then be unbalanced, and will cause a negative acceleration,
resulting in a decrease in speed. But, as the speed decreases, drag will also
decrease. Thus the rate of decrease in speed also decreases.
Newton's first law states: "A body in a state of rest remains at rest, and a body in
motion remains in uniform motion, unless acted upon by some outside force." This
means that if an object is in motion, it will continue in the same direction and at
the same speed until some unbalanced force is applied. And, whenever there are
unbalanced forces acting on an object, that object must change its state of motion.
Newton's second law states: "The rate of change in momentum of an object is
proportional to the force acting on the object, and in the direction of the force." The
momentum of an object may be defined as the force that object would exert to
resist any change of its motion. Newton's third law states: "To every action there is
an equal and opposite reaction." This law means that when a force is applied to
any object, there must be a reaction opposite to and equal to the applied force. If
an object is in motion, and we try to change either the direction or rate of that
motion, the object will exert an equal and opposite force. That force is directly
proportional to the mass of the object, and to the change in its velocity.

Based on the Newton's laws the first chapter is concerned with the derivation of
the mathematical dynamical model of a missile in flight. Also it concerns the
representation of aerodynamic data in the form of force and moment coefficients,
stability derivatives and the effects of atmospheric properties and of airflow
parameters on the aerodynamic forces and moments.

A guided missile may be defined as an unmanned projectile that carries its own
flight control equipment. In addition, the missile carries an explosives payload. The
purpose of a guidance system is to control the path of the missile while it is in
flight. This makes it possible for personnel at ground or mobile launching sites to
hit a desired target, regardless of whether that target is fixed or moving, and
regardless of whether or not it takes deliberate evasive action. The guidance
function may be based on information provided by sources inside the missile, or
on information sent from fixed or mobile control points, or both.

Now we will list several types of guidance systems, such as command guidance,
navigation guidance systems, beam-rider guidance and composite systems. The
term command is used to describe a guidance method in which all guidance
instructions, or commands, come from sources outside the missile. To receive the
commands, the missile contains a receiver that is capable of receiving instructions
from ground stations or from another aircraft. The missile receiver then converts
these commands to guidance information, which is fed to the sections following the
sensor unit. The use of radio for command guidance of high-speed missiles makes
it necessary to use a transmitter that can do more than send simple ON-OFF
pulses. Otherwise, a separate transmitter would be required for each control
function. This would require several radio channels for each missile.

When targets are located at great distances from the launching site, some form of
navigational guidance must be used. Accuracy at long distances is achieved only
after exacting and comprehensive calculations of the flight path have been made.
The mathematical equation for a navigation problem of this type may contain
factors designed to control the movement of the missile about the three axes-pitch,
roll, and yaw. In addition, the equation may contain factors that take into account
acceleration due to outside forces (tail winds) and the inertia of the missile itself.

Beam-rider guidance system (i.e. line of sight) uses the beam pattern of a highly
directional radar antenna as a track between the missile launching point and the
target. Electronic equipment on the ground modulates the beam in such a way
that electronic equipment in the missile can derive guidance instructions from it.
The fixed components of a beam rider system are usually a target-tracking radar, a
computer, and a guidance radar. It is possible to combine the guidance and
tracking function into one beam.

Homing guidance is especially suitable for use during the terminal phase of the
trajectory. The purpose of the homing system is to obtain guidance information
from the target itself, rather than from some other outside source. A homing
missile uses one of two methods in approaching a target. When the missile flies
directly toward the target at all times, the trajectory is known as a zero bearing or
pursuit approach. The second method of approach to the target is called lead angle
course. It is also known as a constant bearing or collision course.

Parallel Navigation (constant bearing): The missile maintains a constant angle with
the target's path, effectively traveling in parallel to the target. The missile's course
remains unchanged unless the target changes its path. This method is simpler but
less effective for fast or maneuvering targets.

Proportional Navigation: The missile adjusts its flight path proportionally to the
rate of change of the line of sight (LOS) angle to the target. The missile constantly
steers towards the predicted future position of the target. It uses a gain factor
(navigation constant) that multiplies the LOS rate. More effective for intercepting
moving targets and compensating for target maneuvers. Continuously adjusts
trajectory based on the LOS rate. Proportional navigation is widely used in modern
missile guidance systems due to its robustness in various combat scenarios. It
aims to intercept the target by predicting its future position.

The proportional navigation (PN) is the guidance law which implements parallel
navigation, but it kept the line-of-sight rate to be zero rather than of constant
direction. PN can be seen as achieving a form of parallel navigation relative to the
moving target, as the missile's path continuously adjusts to maintain a direct
interception course. However, it fundamentally differs from maintaining a constant
direction or angle as in traditional parallel navigation.

Missile trajectories include many types of curves. The exact nature of the curve is
determined by the type of guidance and the nature of the control system used. For
some missiles, the desired trajectory is chosen before the missile is designed, and
the missile is closely limited to that trajectory. Other missiles, such as Regulus,
may offer a choice of trajectories.
Pursuit Curve: Some homing missiles, and some beam riders, follow a pursuit
curve. At any given instant, the course of the missile is directly toward the target.
If missile and target are approaching head-on, or if the missile is engaged in a tail
chase, the pursuit curve may be a straight line unless the target changes course.
But a missile that pursues a crossing target must follow a curved trajectory. As the
missile approaches crossing target, the target bearing rate increases, and the
curvature of the missile course increases correspondingly.

Lead Angle Course: Some homing missiles follow a modified pursuit course. The
deflection of the missile control surfaces is made proportional to the target bearing
rate. The missile flies not toward the target, but toward a point in front of it. The
missile thus develops a lead angle, and the curvature of its course is decreased.

Beam-Rider Trajectory: As we will explain later in the second chapter, a beam-rider


missile may follow either a pursuit curve or a lead-angle course, depending on the
type of system used.

Surface-to-air missile systems are developed to meet specified operational


requirements. In a broad sense these requirements include the size of the defended
area and lethality. In addition, the conditions under which the missile system is to
operate are specified to include the environment and characteristics of the threat
(target). The defended area and threat characteristics determine the missile range
and altitude requirements. The speed and maneuverability of the target influence
the speed and maneuverability required of the missile. The required lethality,
generally expressed as kill probability, translates to requirements for missile
guidance accuracy, dynamic airframe maneuver characteristics, counter-measures
capability, fuzing and warhead characteristics. The kill probability requirements are
usually stated as the probability of achieving specific levels of damage to the target
under specified engagement conditions. One of the principal objectives of modeling
missile flight is to predict how close the missile will approach the target under
varying dynamic and environmental conditions. Miss-distance is often used as a
measure of missile system performance. In general, the smaller the miss distance,
the greater the probability of killing the target. The mathematical analysis of
missile flight is complex and involves nonlinearities, logic sequences, singular
events, and interactions among multiple subsystems. Computer simulation
techniques are ideally suited to this task.

A missile flight simulation is a computational tool that calculates the flight path
and other important parameters of a missile as it leaves the launcher and engages
a target. A simulation is based on mathematical models of the missile, target and
environment, and these mathematical models consist of equations that describe
physical laws and logical sequences. The missile model includes factors such as
missile mass, thrust aerodynamics, guidance and control, and the equations
necessary to calculate the missile attitude and flight path. The target model is
often less detailed but includes sufficient data and equations to determine the
target flight path, signature and countermeasures. The model of the environment
contains, at a minimum, the atmospheric characteristics and gravity. Clouds, sun
position, and terrain or sea surface characteristics are included if they are
important to the purpose of the simulation. Sometimes bread-boarded components
or actual missile hardware is used instead of mathematical models of certain
missile subsystems.

The need for real-time computation is usually the result of using actual missile
hardware in the simulation, which, of course, must run in real time. In this case
the physical simulation consists of lines of instruction for the digital portion, wired
patch-boards for the analog portion, and the actual hardware components (for e.g.
the seeker). The equipment needed to run a hybrid simulation that includes actual
seeker hardware is a digital computer, an analog computer, and a seeker scene
generator. Less complex simulations may require only a digital computer.

In this work, we are interested in investigating a particular guidance problem


related to the minimization of the miss distance and the time of closest approach
of a generic surface-to-air missile. To approach towards the solution of such a
problem, we designed and applied more than five different guidance laws.

The second chapter describes the important functions and concepts in missile
guidance and control system, such as the various types of missile guidance
techniques, the different configurations of control system, autopilot, and their
models and also the different guidance laws applied in the simulation.

The third chapter presents the simulations where the described missile model is
simulated and tested against a moving target with different maneuvers using the
different guidance laws cited in the previous chapter. Then, a qualitative and
comparative study between the results is established.
Dynamic Modeling of • Introduction (Missile Description)
• Missile Components
Surface-Air Missiles • Autopilot System
• Control System
(Missile System Description) • Dynamic Model of a Missile

A missile is defined as a space-traversing unmanned vehicle


which contains the means for controlling its flight path. A guided missile is
considered to operate only above the surface of the Earth, and can be controlled in
flight till interception to achieve destruction of the target. The missiles are
classified by the physical areas of launching and the physical areas containing the
target. The four general categories of the guided missiles are:
• Surface-to-air • Air-to-surface • Surface-to-surface • Air-to-air
Surface-to-air missile systems are designed to meet specified operational
requirements. The purpose of a surface-to-air missile system is to destroy
threatening airborne targets. The system includes the missile flight vehicle and
supportive equipment such as a launcher, any ground-based missile and/or target
trackers, and any ground-based guidance processors.
Guided missiles may also be classified as strategic or tactical, with further
subdivisions depending on the role. Strategic missiles are large missiles, often with
nuclear warheads and very long ranges, meant to destroy the enemy’s ability to
wage war. Tactical missiles, on the other hand, are meant for battlefield use for the
limited purpose of winning the battle or encounter. These can be of different kinds,
depending on their roles.

A guided missile is typically divided into four


subsystems: the airframe, guidance, motor (or propulsion), and warhead.

1 Radom 9 Nozzle
2 Planar array active radar antenna 10 Rear detection antenna
3 Proximity fuze antenna (1/4 spaced at 90°) 11 Hydraulic power unit
4 Warhead 12 Autopilot
5 Fuzing unit 13 Electric converter
6 Fixed wings 14 Rocket motor
7 Umbilical connector 15 Guidance section
8 Moving control fins

A missile seeker is composed of a seeker head to


collect and detect energy from the target, a tracking function to keep the seeker
boresight axis pointed toward the target and a processing function to extract useful
information from the detection and tracking circuits. The seeker usually is
mounted in the nose of the missile where it can have an unobstructed view ahead.
The seeker antenna or optical system is usually mounted on gimbals to permit its
central viewing direction (boresight axis) to be rotated in both azimuth and
elevation relative to the missile centerline.
Radio Frequency Seekers Optical seekers
Passive RF seeker Contrast seeker
Semi-active seeker Laser seeker
Active seeker Cameras seeker
The autopilot in a missile serves as a "translator" between the guidance
processor and the control system.
A system that serves to maintain attitude stability of the missile
and to correct deflections.

A system which evaluates flight information, correlates it with


target data, determines the desired flight path of a missile, and communicates the
necessary commands to the missile flight control system.

is designed to inflict any of several possible kinds of damage on the


enemy. The warhead is the reason that the missile exists, the other components
are intended merely to ensure that the warhead will reach its destination.

The types of warheads that might be used with guided missiles include: external
blast, fragmentation, shaped-charge, explosive-pellet, chemical, biological, nuclear,
continuous rod, clustered, thermal, illuminating, psychological, and dummy.

The Airframe is the cylindrical tube structure that carries the warhead
to the target, and houses all the missile subsystems, attached end-to-end, and
supports the control fins, stabilizing fins, and wings (if any). The method of control
influences the airframe configuration. Configurations with canard control, tail
control, and wing control. Airframe deflection (aero-elastic effect) is an important
consideration in missile design.

provides the energy required to move the missile from the


launcher to the target. There are two basic types of jet propulsion power plants
used in missile propulsion systems atmospheric jet and thermal jet.

Atmospheric jet propulsion system: Any jet-propelled system that obtains oxygen
from the surrounding atmosphere to support the combustion of its fuel is an
atmospheric jet engine.
Pulsejet
Ramjet
Turbojet
Thermal jet propulsion system: Thermal jets include solid propellant, liquid
propellant, and combined propellant systems.
Liquid propellant
Solid propellant
Combined propellant

Missile models are based on mathematical


equations that describe the dynamic motions of missiles that result from the forces
and moments acting upon them. The mathematical tools employed are the
equations of motion, which describe the relationships between the forces acting on
the missile and the resulting missile motion. Three-degree-of-freedom models
employ translational equations of motion; six-degree-of-freedom models employ, in
addition, rotational equations of motion. The inputs to the equations of motion are
the forces and moments acting on the missile; the outputs are the missile
accelerations that result from the applied forces and moments.
Surface-to-air missile systems are designed to meet specified operational
requirements. The purpose of a surface-to-air missile system is to destroy
threatening airborne targets. The system includes the missile flight vehicle and
supportive equipment such as a launcher, any ground-based missile and/or target
trackers, and any ground-based guidance processors. As the tracking system
continues to measure relative motion, a guidance processor derives missile
maneuver commands to guide the missile to intercept the target. The maneuver
commands are transformed into missile control-surface deflection commands by
an autopilot and a control system supplies the actuator power to rotate the control
surfaces. Aerodynamic lift on the missile generated by control-surface deflections
produces maneuvers that are responsive to guidance commands.

A missile flight simulation is a computational tool that calculates the flight path
and other important parameters of a missile as it leaves the launcher and engages
a target. A simulation is based on mathematical models of the missile, target and
environment, and these mathematical models consist of equations that describe
physical laws and logical sequences. The missile model includes factors such as
missile mass, thrust aerodynamics, guidance and control, and the equations
necessary to calculate the missile attitude and flight path. The target model is
often less detailed but includes sufficient data and equations to determine the
target flight path, signature, and countermeasures. The model of the environment
contains, at a minimum, the atmospheric characteristics and gravity. Clouds,
‘haze, sun position, and terrain or sea surface characteristics are included if they
are important to the purpose of the simulation. Sometimes actual missile
hardware is used instead of mathematical models of certain missile subsystems.

The physical laws in the simulation are those governing the motion of the missile
and target and those affecting any simulated subsystems. For example, the
equations of motion of the missile determine the acceleration, velocity, and
position resulting from the forces due to gravity, thrust, and aerodynamics. Other
equations governing physical processes may be required to simulate subsystems
such as the target tracking system or the missile control system.

As pointed out 𝐅Total (𝑡) = 𝑚. 𝐯̇ (𝑡), was originally derived


on the assumption of constant mass, but this same equation is obtained also when
the effects of mass variation due to propellant burning are taken into account. A
frequent error in the application of Newton’s equations of motion to systems with
variable mass is to assume that the rate of change of linear momentum is given by
𝑑(𝑚. 𝐯) 𝑑𝑚 𝑑𝐯 (𝐖𝐫𝐨𝐧𝐠)
𝐅Total (𝑡) = = 𝐯(𝑡). + 𝑚(𝑡). [N = kg. m/s]
𝑑𝑡 𝑑𝑡 𝑑𝑡
Where: 𝐅Total = vector sum of forces acting on the missile, [𝑁]. 𝑡= time, [𝑠]
𝑚(𝑡) = instantaneous mass of missile (includes mass of unburned fuel), [kg]
𝑚̇(𝑡) = rate of change of missile mass 𝑚(𝑡) (𝑚̇𝑒 (𝑡) = −𝑚̇(𝑡)), [kg/s]
𝐯(𝑡) = absolute linear velocity vector of missile, [𝑚/s]
𝐯̇ (𝑡) = absolute acceleration vector of center of mass of missile, [𝑚/s 2 ].
The correct rate of change of momentum of the system must take into account the
fact that not all mass particles in the system have the same velocity. In the case of
a missile, the missile itself (including unburned propellant) has absolute velocity
𝐯(𝑡), and the exhaust gases have absolute velocity 𝐯𝑒 (𝑡) and relative velocity
𝐯𝑟𝑒 (𝑡) = 𝐯𝑒 (𝑡) − 𝐯(𝑡) with respect to the center of mass of the missile.

For a missile the external forces 𝐅ext (𝑡) consist of aerodynamic forces, the pressure
component of thrust, and gravity. These external forces are applied directly to the
missile body; therefore, they affect only the portion of the total system momentum
attributable to the missile. From the definition of linear momentum an appropriate
result for the time interval Δ𝑡 can be written as: 𝐅ext = Δ𝐋/Δ𝑡 = (𝐋f − 𝐋𝑖 )/Δ𝑡 [N]
where 𝑳𝑖 = initial total momentum at beginning of time interval, [𝑁. s]. 𝐋f = final
total system momentum at end of time interval, [𝑁. s].

The values of the momentum of the total system at the beginning and end of the
time interval are given by 𝐋𝑖 = 𝑚𝐯 [N. s], and 𝐋f = (𝑚 − Δ𝑚𝑒 )(𝐯 + Δ𝐯) + Δ𝑚𝑒 𝐯𝑒 [N. s]
where: 𝑚 = missile mass at the beginning, [kg]. Δ𝑚𝑒 = mass of exhaust gases
expelled from missile during time interval, [kg]. 𝐯(𝑡) = absolute velocity of missile at
beginning of time interval, [m/s]. Δ𝐯 = change in missile velocity during time
interval, [m/s]. 𝐯𝑒 (𝑡) = absolute velocity of exhaust gases, [m/s]. This will result to
the following compact form equation:
𝐅ext (𝑡) = (𝐋f − 𝐋𝑖 )/Δ𝑡 = [(𝑚 − Δ𝑚𝑒 )(𝐯 + Δ𝐯) + Δ𝑚𝑒 𝐯𝑒 − 𝑚𝐯]/Δ𝑡 [N]

If Δ𝑡 approaches zero, i.e., Δ𝑡 → 0, then 𝐯̇ ≈ Δ𝐯/Δ𝑡, 𝑚𝑒 ≈ Δ𝑚𝑒 /Δ𝑡 and Δ𝑚𝑒 Δ𝐯 ≈ 𝟎.


This will lead to the next result 𝐅ext (𝑡) = 𝑚𝐯̇ + 𝑚̇𝑒 (𝐯𝑒 − 𝐯) = 𝑚𝐯̇ + 𝑚̇𝑒 𝐯𝑟𝑒 [N]. This is
known as 𝑀𝑒𝑠ℎ𝑐ℎ𝑒𝑟𝑠𝑘𝑦 𝑒𝑞𝑢𝑎𝑡𝑖𝑜𝑛 for variable mass system.

The total thrust vector 𝐅𝑃 (𝑡) (i.e. propulsive forces) produced by a jet motor at the
end of missile is composed of two parts: the momentum thrust and the pressure
thrust, so 𝐅𝑃 (𝑡) = 𝐅mom + 𝐅pres . The portion of the total thrust attributed to this
momentum change has magnitude 𝐹mom = 𝑚̇𝑒 ‖𝐯𝑟𝑒 ‖ where 𝑚̇𝑒 is the mass rate of
flow of the exhaust gases and (𝐯𝑟𝑒 (𝑡) = 𝐯𝑒 (𝑡) − 𝐯(𝑡)) is the velocity of the exhaust
gases relative to the missile, 𝐯 is the missile velocity.

The average pressure 𝑝𝑒 of the expanding exhaust gases at the exit plane of the jet
nozzle acts over the exit area 𝐴𝑒 of the jet nozzle. The remainder of the missile is
surrounded by the ambient atmospheric pressure 𝑝𝑎 . This imbalance of pressure
constitutes the pressure thrust, which has magnitude 𝐹pres = (𝑝𝑒 − 𝑝𝑎 )𝐴𝑒 .
Combining the two thrust portions in vector form, the total thrust force on the
missile is given by: 𝐅𝑃 (𝑡) = 𝐅mom + 𝐅pres = −[𝑚̇𝑒 𝐯𝑟𝑒 + (𝑝𝑒 − 𝑝𝑎 )𝐴𝑒 𝐮𝑣𝑒 ]
𝐴𝑒 = jet nozzle exit area, [m2 ]. 𝐅𝑃 (𝑡) = total instantaneous thrust force vector, [N].
𝑝𝑎 = ambient atmospheric pressure, [Pa]. 𝑝𝑒 = the average pressure across 𝐴𝑒 , [Pa].
𝐮𝑣𝑒 = unit vector in direction of relative exhaust velocity. 𝐯𝑟𝑒 = velocity vector of
exhaust gas relative to center of mass of missile, [m/s].

As stated before the vector sum of forces external to the total, closed system
𝐅ext (𝑡) = 𝑚(𝑡). 𝐯̇ (𝑡) + 𝑚̇𝑒 (𝑡). 𝐯𝑟𝑒 (𝑡) consists of the aerodynamic force 𝐅𝐴 , the pressure
force 𝐅pres = −(𝑝𝑒 − 𝑝𝑎 )𝐴𝑒 (𝐮𝑣𝑒 ), and the gravitational force 𝐅g .

𝐅ext = 𝐅𝐴 + 𝐅g + 𝐅pres = 𝑚𝐯̇ + 𝑚̇𝑒 𝐯𝑟𝑒 ⟹ 𝐅𝐴 + 𝐅g − [𝑚̇𝑒 𝐯𝑟𝑒 + (𝑝𝑒 − 𝑝𝑎 )𝐴𝑒 (𝐮𝑣𝑒 )] = 𝑚𝐯̇ [N]

Substituting the definition of 𝐅𝑃 , which includes both the pressure and momentum
components of thrust, into this last equation, gives: 𝐅𝐴 + 𝐅g + 𝐅𝑃 = 𝑚(𝑡). 𝐯̇ (𝑡) [N].
Finally, setting 𝐅Total (𝑡) = 𝐅𝐴 + 𝐅g + 𝐅𝑃 acting directly on the missile allows one to
write: 𝐅Total (𝑡) = 𝑚(𝑡). 𝐯̇ [N]. This is the familiar form of Newton’s equation. Thus
it is shown that a missile with a jet motor is analyzed in the same way as any
problem having constant mass except that 𝑚 to be used is a function of time.

The general motion of a body can be described by the


translation of the center of mass and the rotation of the body about its center of
mass. In order to solve the newton’s second low, one must know 𝑟(𝑡) (the vector
position) for all the particles constituting the body that describing the relative
motion of the particles with respect to the center of mass, be obtained by
integrating additional differential equations in time. Such equations of relative
motion are obtained by taking into account the internal forces acting on the
individual particles. However, if the body is rigid, the relative distances of all
particles with respect to the center of mass are fixed. The translational and
rotational motions can be studied separately, provided the net force and moment
vectors do not depend upon the rotational and translational motions, respectively.
For an atmospheric flight vehicle, the aerodynamic force depends upon the
vehicle’s attitude (rotational variables), and aerodynamic and thrust moments
depend upon the speed and altitude (translational variables); thus, the two
motions are inherently coupled. We can summarize it in this statement:
𝐆𝐞𝐧𝐞𝐫𝐚𝐥 𝐦𝐨𝐭𝐢𝐨𝐧 = 𝐓𝐫𝐚𝐧𝐬𝐥𝐚𝐭𝐢𝐨𝐧 𝐦𝐨𝐭𝐢𝐨𝐧 + 𝐑𝐨𝐭𝐚𝐭𝐢𝐨𝐧𝐚𝐥 𝐦𝐨𝐭𝐢𝐨𝐧

For a rotating continuum rigid body, the relation 𝐯 = 𝛚 × 𝐫 (where: 𝐫 is the position
vector of the center of mass "radial from the rotation axis" and 𝐯 is the tangential
velocity of the particle) holds for each point in the rigid body. If 𝐑(𝑡) is a rotation
matrix that rotates 𝐫0 = const to be 𝐫(𝑡) = 𝐑(𝑡)𝐫0 , and we define the tensor of
rotation 𝐖 = 𝐑𝐑̇𝑇 = −𝐑̇𝐑𝑇 (i.e. 𝐑𝐑𝑇 = 𝑰3 ⇒ 𝐑̇𝐑𝑇 + 𝐑𝐑̇𝑇 = 𝟎) then
𝑑𝐫 𝑑𝐑 𝑑𝐑
𝐯= = 𝐫0 = { 𝐑𝑇 } 𝐫 = 𝐖𝐫 = 𝛚 × 𝐫 with 𝐖 𝑇 = −𝐖 and 𝐖 = [𝛚]
𝑑𝑡 𝑑𝑡 𝑑𝑡

With 𝛚 is a vector formed by skew-symmetric tensor 𝐖, and is called angular


velocity. So it is always valid that: 𝐯 = 𝛚 × 𝐫 ⟺ 𝛚 = 𝐯 × 𝐫/(𝐫. 𝐫) and 𝐚𝑛 = 𝛚 × 𝐯.
In order to calculate the derivative of some rotating unit vectors 𝐢, 𝐣, and 𝐤 we
assume that 𝐫 = 𝐢 is the radius of a point 𝑀 on the axis 𝑥 at unit distance from the
origin. Then 𝐯 = 𝑑𝐫/𝑑𝑡 = 𝛚 × 𝐫 ⟺ 𝑑𝐢/𝑑𝑡 = 𝝎 × 𝐢. Similar relationships are obtained
for the derivatives of 𝐣, and 𝐤, and finally we obtain the equations 𝑑𝐢/𝑑𝑡 = 𝛚 × 𝐢,
𝑑𝐣/𝑑𝑡 = 𝛚 × 𝐣, and 𝑑𝐤/𝑑𝑡 = 𝛚 × 𝐤 which are known as the 𝐏𝐨𝐢𝐬𝐬𝐨𝐧 𝐞𝐪𝐮𝐚𝐭𝐢𝐨𝐧𝐬.
Consider a fixed system (O 𝑥𝐸 𝑦𝐸 𝑧𝐸 ), assumed to be the inertial frame and another
rotating system (O 𝑥𝐵 𝑦𝐵 𝑧𝐵 ), assumed to be the body’s fixed frame, which rotate
with respect to the first one by an angular velocity 𝛚. Let 𝐢, 𝐣, and 𝐤 be a unit
vectors along the axes of the rotating system. Let 𝐯 be an arbitrary vector with
components v𝑥 , v𝑦 and v𝑧 along the rotating axes. Then:
𝑑𝐯 𝑑v𝑥 𝑑v𝑦 𝑑v𝑧 𝑑𝐢 𝑑𝐣 𝑑𝐤 So we deduce that:
( ) =( 𝐢+ 𝐣+𝐤 ) + v𝑥 + v𝑦 + v𝑧
𝑑𝑡 𝐸 𝑑𝑡 𝑑𝑡 𝑑𝑡 𝑑𝑡 𝑑𝑡 𝑑𝑡 | 𝑑𝐯 𝑑𝐯
𝑑𝐯 𝑑𝐢 𝑑𝐣 𝑑𝐤 ( ) =( ) +𝛚×𝐯
= ( ) + v𝑥 + v𝑦 + v𝑧 𝑑𝑡 𝐸 𝑑𝑡 𝐵
𝑑𝑡 𝐵 𝑑𝑡 𝑑𝑡 𝑑𝑡
|
𝑑𝐯
= ( ) + 𝛚 × (v𝑥 𝐢 + v𝑦 𝐣 + v𝑧 𝐤) 𝑑𝐇 𝑑𝐇
𝑑𝑡 𝐵 ( ) =( ) +𝛚×𝐇
𝑑𝐯 | 𝑑𝑡 𝐸 𝑑𝑡 𝐵
=( ) +𝛚×𝐯 where 𝐇 is the angular momentum
𝑑𝑡 𝐵
If we consider a missile to be rigid body which is constituting by a collection of
particles of elemental mass 𝛿𝑚, and 𝐯(𝑡) the velocity of
the elemental mass relative to the inertial frame, and 𝛿𝐅
the resulting force acting on the element mass. So from 𝛿𝑚 𝛿𝐅⃗
Newton’s second law we have: 𝛿𝐅 = 𝛿𝑚. (𝑑𝐯/𝑑𝑡). The total
external force acting on the missile is found by summing 𝐯𝒐
𝐫⃗
all the elements of the missile. Therefore, 𝐅 = ∑ 𝛿𝐅 The 𝐜𝐦
velocity of the differential mass 𝛿𝑚 is: 𝐯(𝑡) = 𝐯𝑜 (𝑡) + 𝑑𝐫/𝑑𝑡
with 𝐯𝑜 (𝑡) is the velocity of the center of mass of the Figure. velocity of 𝛿𝑚
missile and 𝑑𝐫/𝑑𝑡 is the velocity of the element relative to center of mass.
𝑑 𝑑𝐫 𝑑𝐯𝑜 𝑑 𝑑𝐫 𝑑𝐯𝑜 𝑑 2
𝐅 = ∑ 𝛿𝐅 = ∑ {𝐯𝑜 + } . 𝛿𝑚 = ∑ 𝛿𝑚 + ∑ 𝛿𝑚 = 𝑚 + 2 ∑ 𝐫𝛿𝑚
𝑑𝑡 𝑑𝑡 𝑑𝑡 𝑑𝑡 𝑑𝑡 𝑑𝑡 𝑑𝑡
Because 𝐫 is measured from the center of mass then ∑ 𝐫𝛿𝑚 = 𝟎 and therefore we
get the following well-known equation: 𝐅 = 𝑚. (𝑑𝐯𝑜 /𝑑𝑡).

Similarly, we can develop the moment equation referred to a moving center of


mass. We know that 𝐌 = 𝑑𝐇/𝑑𝑡 = 𝕀𝒐 . (𝑑𝛚/𝑑𝑡) where 𝐇 = 𝕀𝒐 . 𝛚 is the angular
momentum, and 𝐌 is the moment (or torque). For the differential element of mass
𝛿𝑚, the moment equation can be written as: 𝐌 = ∑ 𝛿𝐌 = 𝑑𝐇/𝑑𝑡 where 𝐇 = ∑ 𝛿𝐇 and
𝛿𝐇 = 𝐫 × 𝐯. 𝛿𝑚 so we have

𝐇 = ∑ 𝛿𝐇 = ∑ 𝐫 × [𝐯𝑜 + 𝛚 × 𝐫]. 𝛿𝑚 = ∑ 𝐫 × 𝐯𝑜 . 𝛿𝑚 + ∑[𝐫 × (𝛚 × 𝐫)]. 𝛿𝑚

we know that ∑ 𝐫𝛿𝑚 = 𝟎 then: 𝑯 = ∑[𝐫 × (𝛚 × 𝐫)]. 𝛿𝑚, hence 𝐇 = 𝕀𝒐 . 𝛚 where 𝕀𝒐 is


called the inertia matrix (or the inertia tensor). Later on we can prove that this
inertia matrix around the center of mass is given by 𝕀𝒐 = ∑[𝛀𝑟 ][𝛀𝑟 ]𝑇 . 𝛿𝑚.
Mathematically, Newton’s second law can be expressed in terms of conservation of
both linear and angular momentum by the following vector equations:
𝑑(𝑚𝐯) 𝑑𝐯𝑜 𝑑𝐇 𝑑 𝑑𝛚
∑𝐅 = { } = 𝑚. { } ; and ∑𝐌 = { } = { (𝕀𝒐 . 𝛚)} = 𝕀𝒐 . { }
𝑑𝑡 𝐸 𝑑𝑡 𝐸 𝑑𝑡 𝐸 𝑑𝑡 𝐸 𝑑𝑡 𝐸
Note: In the next subsection we use the notion 𝐯𝑚 instead of 𝐯𝑜 which is the
missile velocity measured at its center of mass.

It is well known that on each material body


there are acting forces and moments. Newton's law, concerning an inertial system
states that: 𝐅Total = 𝑚(𝑡). {𝐚𝑚 (𝑡)}𝐸 . It is also known that in the case of a kinematic
system, the following relation is valid: (𝑑𝐱/𝑑𝑡)𝐸 = (𝑑𝐱/𝑑𝑡)𝐵 + 𝛚 × (𝐱). Thus in case of
a freely maneuvering body, we have
𝑑𝐯𝑚 𝑑𝐯𝑚
𝐅Total (𝑡) = 𝑚(𝑡). ( ) = 𝑚(𝑡). [( ) + 𝛚(𝑡) × 𝐯𝑚 (𝑡)]
𝑑𝑡 𝐸 𝑑𝑡 𝐵
where: 𝐅Total (𝑡): total force acting on a missile (thrust, aerodynamic, gravitational),
𝐯𝑚 (𝑡): vector velocity of a missile, 𝛚(𝑡): angular rate of a missile, 𝑚 = f(𝑡) [kg] is
the total mass of missile (i.e. it is a function of time), 𝐚𝑚 = 𝑑𝐯𝑚 /𝑑𝑡 [m/s2 ]: linear
acceleration, and 𝜶𝑚 = 𝛚 × 𝐯𝑚 [rad/s2 ]: centripetal acceleration.

The total force acting on a missile is given by: 𝐅Total = 𝐅G + 𝐅A + 𝐅P where: 𝐅G : missile
weight, 𝐅P : vector thrust of missile motor (i.e. Propulsive forces) and 𝐅A : missile
vector aerodynamic force. The forces 𝐅P and 𝐅A are controllable applied to the
missile forces while 𝐅G = 𝑚𝐠 is uncontrollable existing force.

In this study, the missile is considered as moving into the perceivable three-
dimensional (3𝐷) space. Thus, making use of international notation, the following
expressions are derived:
𝐯𝑚 = 𝑢𝐢𝒃 + 𝑣𝐣𝒃 + 𝑤𝐤 𝒃 𝑑𝐯𝑚 𝑑𝑢 𝑑𝑣 𝑑𝑤
( ) = ( ) 𝐢𝒃 + ( ) 𝐣 𝒃 + ( ) 𝐤 𝒃
𝛚 = 𝑝𝐢𝒃 + 𝑞𝐣𝒃 + 𝑟𝐤 𝒃 𝑑𝑡 𝐵 𝑑𝑡 𝑑𝑡 𝑑𝑡
| 𝐅Total = 𝐹x 𝐢𝒃 + 𝐹y 𝐣𝒃 + 𝐹z 𝐤 𝒃
𝐅G = 𝐹gx 𝐢𝒃 + 𝐹gy 𝐣𝒃 + 𝐹gz 𝐤 𝒃 𝐹x = 𝐹gx + 𝐹Px + 𝐹Ax = 𝑋 + 𝑚g x
𝐅P = 𝐹Px 𝐢𝒃 + 𝐹Py 𝐣𝒃 + 𝐹Pz 𝐤 𝒃 |
𝐹y = 𝐹gy + 𝐹Py + 𝐹Ay = 𝑌 + 𝑚g y
𝐅A = 𝐹Ax 𝐢𝒃 + 𝐹Ay 𝐣𝒃 + 𝐹Az 𝐤 𝒃 𝐹z = 𝐹gz + 𝐹Pz + 𝐹Az = 𝑍 + 𝑚g z
𝐢𝒃 𝐣𝒃 𝐤𝒃
𝛚 × 𝐯𝑚 = det [ 𝑝 𝑞 𝑟] = (𝑞𝑤 − 𝑟𝑣)𝐢𝒃 + (𝑟𝑢 − 𝑝𝑤)𝐣𝒃 + (𝑝𝑣 − 𝑞𝑢)𝐤 𝒃
𝑢 𝑣 𝑤
𝐯𝑚 = 𝑢𝐢𝒃 + 𝑣𝐣𝒃 + w𝐤 𝒃 , [m/s] where 𝑢, 𝑣, 𝑤: are components of absolute linear velocity
vector 𝐯𝑚 expressed in body coordinate system. (𝐢𝒃 , 𝐣𝒃 , 𝐤 𝒃 ) are the unit vectors along
the respective missile body axes.
𝐚𝑚 = 𝑢̇ 𝐢𝒃 + 𝑣̇ 𝐣𝒃 + ẇ𝐤 𝒃 , [m/s2 ] where 𝑢̇ , 𝑣̇ , 𝑤̇ : are components of linear (translational)
acceleration expressed in the body fixed frame. 𝑢̇ : longitudinal acceleration, 𝑣̇ :
lateral acceleration and 𝑤̇ : vertical acceleration.
𝝎 = 𝑝𝐢𝒃 + 𝑞𝐣𝒃 + 𝑟𝐤 𝒃 , [rad/s] where 𝑝, 𝑞, 𝑟: are components of the angular velocity
vector expressed in body fixed frame (along the roll, pitch, and yaw, respectively).
𝜶𝑚 = 𝑝̇ 𝐢𝒃 + 𝑞̇ 𝐣𝒃 + 𝑟̇ 𝐤 𝒃 , [rad/s 2 ] where 𝑝̇ , 𝑞̇ , 𝑟̇ : are components of angular acceleration
expressed in body coordinate system (along the roll, pitch, and yaw, respectively).
𝑇
𝐅Total = [F𝑥 (𝑡), F𝑦 (𝑡), F𝑧 (𝑡)] : is the external force includes the sum of aerodynamic,
pressure thrust, and gravitational forces (expressed in body coordinate system).

𝐹𝑥 (𝑡) = 𝐹𝐴𝑥 + 𝐹𝑃𝑥 + 𝐹g𝑥 ; 𝐹𝑦 (𝑡) = 𝐹𝐴𝑦 + 𝐹𝑃𝑦 + 𝐹g𝑦 ; and 𝐹𝑧 (𝑡) = 𝐹𝐴𝑧 + 𝐹𝑃𝑧 + 𝐹g𝑧
Notice that: 𝐹Ax : drag force, 𝐹Ay : side force and 𝐹Az : lift force. Equating the
components of 𝐅Total = 𝑚[𝐚𝐵 + 𝛚 × 𝐯𝑚 ] yields the missile’s linear equations of motion
𝑑𝑢 𝑑𝑣 𝑑𝑤
𝐹x (𝑡) = 𝑚 [ + 𝑞𝑤 − 𝑟𝑣] ; 𝐹y (𝑡) = 𝑚 [ + 𝑟𝑢 − 𝑝𝑤] ; 𝐹z (𝑡) = 𝑚 [ + 𝑝𝑣 − 𝑞𝑢]
𝑑𝑡 𝑑𝑡 𝑑𝑡
Substituting the components of 𝐹x , 𝐹y , and 𝐹z and rearrange the equations using the
necessery operations we obtain the final translation equations of missile model

𝑑𝑢/𝑑𝑡 = [𝐹gx + 𝐹Px + 𝐹Ax ]/𝑚 − (𝑞𝑤 − 𝑟𝑣)


𝑑𝐯𝑚
𝐅Total (𝑡) = 𝑚 [( ) + 𝛚 × 𝐯𝑚 ] ⟺ 𝑑𝑣/𝑑𝑡 = [𝐹gy + 𝐹Py + 𝐹Ay ]/𝑚 − (𝑟𝑢 − 𝑝𝑤)
𝑑𝑡 𝐵
𝑑𝑤/𝑑𝑡 = [𝐹gz + 𝐹Pz + 𝐹Az ]/𝑚 − (𝑝𝑣 − 𝑞𝑢)

The rotational motion of a vehicle is important for


various reasons (aerodynamics, pointing of weapons, payload, or antennas, etc.)
and governs the instantaneous attitude (orientation). It was evident that the
instantaneous attitude depends not only upon the rotational kinematics, but also
on rotational dynamics which determine how the attitude parameters change with
time for a specified angular velocity.

• It is well known that the absolute rate of change of the moment of momentum
(with respect to flight path axis), for a freely maneuvering body in 3𝐷, is given by:

𝑑𝐇 𝑑𝐇 𝐌 = 𝐿𝐢𝒃 + 𝑀𝐣𝒃 + 𝑁𝐤 𝒃
𝐌= ( ) =( ) +𝛚×𝐇 with
𝑑𝑡 𝐸 𝑑𝑡 𝐵 𝐇 = 𝐻𝑥 𝐢𝒃 + 𝐻𝑦 𝐣𝒃 + 𝐻𝑧 𝐤 𝒃

Before we develop these equations, an expression for the angular momentum 𝐇 is


needed to this end.
where
𝐇 = ∑[𝐫 × (𝛚 × 𝐫)]. 𝛿𝑚
𝐫 = 𝑥𝐢𝒃 + 𝑦𝐣𝒃 + 𝑧𝐤 𝒃
|
= − ∑[𝐫 × (𝐫 × 𝛚)]. 𝛿𝑚 𝝎 = 𝑝𝐢𝒃 + 𝑞𝐣𝒃 + 𝑟𝐤 𝒃
𝐇 = 𝐻𝑥 𝐢𝒃 + 𝐻𝑦 𝐣𝒃 + 𝐻𝑧 𝐤 𝒃
= − ∑[𝐫 × ([𝛀𝑟 ]. 𝛚)]. 𝛿𝑚 0 −𝑧 𝑦
|
[𝛀𝑟 ] = [ 𝑧 0 −𝑥 ]
= ∑[([𝛀𝑟 ]. 𝛚) × 𝐫]. 𝛿𝑚 −𝑦 𝑥 0
𝑇
[𝛀𝑟 ] = −[𝛀𝑟 ]
= ∑[𝛀𝑟 ][𝛀𝑟 ]𝑇 . 𝛚. 𝛿𝑚 |
= 𝕀𝒐 . 𝛚 𝕀𝒐 = ∑[𝛀𝑟 ][𝛀𝑟 ]𝑇 . 𝛿𝑚

(𝑦 2 + 𝑧 2 ) −𝑥𝑦 −𝑥𝑧 𝐼𝑥 −𝐼𝑥𝑦 −𝐼𝑥𝑧


𝕀𝒐 = ∑[𝛀𝑟 ][𝛀𝑟 ]𝑇 . 𝛿𝑚 = ∑ [ −𝑥𝑦 2 2
(𝑥 + 𝑧 ) −𝑦𝑧 ] . 𝛿𝑚 = [−𝐼𝑥𝑦 𝐼𝑦 −𝐼𝑦𝑧 ]
−𝑥𝑧 −𝑦𝑧 (𝑥 + 𝑦 2 )
2
−𝐼𝑥𝑧 −𝐼𝑦𝑧 𝐼𝑧
Notice that 𝐇 = 𝕀𝒐 . 𝛚 = (𝑝𝐼𝑥 − 𝑞𝐼𝑦𝑥 − 𝑟𝐼𝑥𝑧 )𝐢𝒃 + (𝑞𝐼𝑦 − 𝑝𝐼𝑦𝑥 − 𝑟𝐼𝑦𝑧 )𝐣𝒃 + (𝑟𝐼𝑧 − 𝑞𝐼𝑦𝑧 − 𝑝𝐼𝑥𝑧 )𝐤 𝒃
combine this with 𝐇 = 𝐻𝑥 𝐢𝒃 + 𝐻𝑦 𝐣𝒃 + 𝐻𝑧 𝐤 𝒃 we get

𝐻𝑥 = (𝑝𝐼𝑥 − 𝑞𝐼𝑦𝑥 − 𝑟𝐼𝑥𝑧 ) = 𝑝(𝑦 2 + 𝑧 2 ) − 𝑞𝑥𝑦 − 𝑟𝑥𝑧 𝐢𝒃 𝐣 𝒃 𝐤 𝒃 𝑞𝐻𝑧 − 𝑟𝐻𝑦


2 2)
𝐻𝑦 = (𝑞𝐼𝑦 − 𝑝𝐼𝑦𝑥 − 𝑟𝐼𝑦𝑧 ) = 𝑞(𝑥 + 𝑧 − 𝑝𝑥𝑦 − 𝑟𝑦𝑧 and 𝛚 × 𝐇 = | 𝑝 𝑞 𝑟 | = [ 𝑟𝐻𝑥 − 𝑝𝐻𝑧 ]
2 2)
𝐻𝑧 = (𝑟𝐼𝑧 − 𝑞𝐼𝑦𝑧 − 𝑝𝐼𝑥𝑧 ) = 𝑟(𝑥 + 𝑦 − 𝑝𝑥𝑧 − 𝑞𝑦𝑧 𝐻 𝐻
𝑥 𝑦 𝑧 𝐻 𝑝𝐻𝑦 − 𝑞𝐻𝑥

• Now, let we try to evaluate the derivative of 𝐇 with respect to time, that is
𝑑𝐇 𝑑 𝑑𝕀𝒐 𝑑𝛚
𝐌= = (𝕀𝒐 . 𝛚) = 𝛚 + 𝕀𝒐
𝑑𝑡 𝑑𝑡 𝑑𝑡 𝑑𝑡
If this equation were evaluated in an inertial reference frame, the moments of
inertia about the frame axes would change as the body experienced rotational
motion and result in nonzero values of 𝕀𝒐 . However, if a reference frame fixed to the
body were employed, the inertia matrix would not be changed by body motion and
thus would provide another motivation to select the body frame. During the
operation of the propulsion system of a missile, there is another source of change
in the moments of inertia that is not related to body motion. As the propellant
mass is expelled from the missile, the moments of inertia change. This change in
the value of 𝕀𝒐 is usually updated continuously in a flight simulation, but 𝑑𝕀𝒐 /𝑑𝑡
the time rate of change 𝕀𝒐 is usually small enough to be neglected in 𝑑𝐇/𝑑𝑡. Thus
selecting the body frame and assuming that the time rates of change of the
moments of inertia caused by propellant expulsion are small cause 𝕀̇𝒐 . 𝛚 to vanish.
If 𝕀𝒐 is calculated relative to the body frame axes and 𝛚 is expressed in the body
frame, the time rate of change of angular momentum relative to that frame (the
first term on 𝐌) is given by
𝐿 𝐼𝑥 −𝐼𝑥𝑦 −𝐼𝑥𝑧 𝑝̇ 𝑞𝐻𝑧 − 𝑟𝐻𝑦
𝑑𝐇 𝑑𝛚
[𝑀] = 𝐌 = ( ) + 𝛚 × 𝐇 = 𝕀𝒐 + 𝛚 × 𝐇 = [−𝐼𝑥𝑦 𝐼𝑦 −𝐼𝑦𝑧 ] [𝑞̇ ] + [ 𝑟𝐻𝑥 − 𝑝𝐻𝑧 ]
𝑑𝑡 𝐵 𝑑𝑡 𝑝𝐻𝑦 − 𝑞𝐻𝑥
𝑁 −𝐼𝑥𝑧 −𝐼𝑦𝑧 𝐼𝑧 𝑟̇
Using the matrix inversion we can write

𝐿 𝑝̇ 𝐼𝑥 −𝐼𝑥𝑦 −𝐼𝑥𝑧 −1 𝐿 − (𝑞𝐻𝑧 − 𝑟𝐻𝑦 )


𝑑𝐇
[𝑀] = ( ) + 𝛚 × 𝐇 ⟺ [𝑞̇ ] = [−𝐼𝑥𝑦 𝐼𝑦 −𝐼𝑦𝑧 ] [ 𝑀 − (𝑟𝐻𝑥 − 𝑝𝐻𝑧 ) ]
𝑑𝑡 𝐵
𝑁 𝑟̇ −𝐼𝑥𝑧 −𝐼𝑦𝑧 𝐼𝑧 𝑁 − (𝑝𝐻𝑦 − 𝑞𝐻𝑥 )
Due to the assumption that the given missile has a cruciform symmetry i.e. the
product of inertia 𝐼𝑥𝑦 = 𝐼𝑦𝑧 = 0 and therefore the inverse of inertia matrix is:
−1 2 ) 2 )
𝐼𝑥 0 −𝐼𝑥𝑧 𝐼𝑧 /(𝐼𝑥 𝐼𝑧 − 𝐼𝑥𝑧 0 𝐼𝑥𝑧 /(𝐼𝑥 𝐼𝑧 − 𝐼𝑥𝑧
[ 0 𝐼𝑦 0 ] =[ 0 1/𝐼𝑦 0 ]
−𝐼𝑥𝑧 0 𝐼𝑧 2 )
𝐼𝑥𝑧 /(𝐼𝑥 𝐼𝑧 − 𝐼𝑥𝑧 0 2 )
𝐼𝑥 /(𝐼𝑥 𝐼𝑧 − 𝐼𝑥𝑧

So the equations can be simplified as:

𝑝̇
2 )
𝐼𝑧 /(𝐼𝑥 𝐼𝑧 − 𝐼𝑥𝑧 0 2 )
𝐼𝑥𝑧 /(𝐼𝑥 𝐼𝑧 − 𝐼𝑥𝑧 𝐿 − (𝑞𝑟𝐼𝑧 − 𝑝𝑞𝐼𝑥𝑧 − 𝑟𝑞𝐼𝑦 )
[𝑞̇ ] = [ 0 1/𝐼𝑦 0 ] [𝑀 − (𝑟𝑝𝐼𝑥 − 𝑟 2 𝐼𝑥𝑧 − 𝑟𝑝𝐼𝑧 + 𝑝2 𝐼𝑥𝑧 )]
𝑟̇ 2
𝐼𝑥𝑧 /(𝐼𝑥 𝐼𝑧 − 𝐼𝑥𝑧 ) 0 2 )
𝐼𝑥 /(𝐼𝑥 𝐼𝑧 − 𝐼𝑥𝑧 𝑁 − (𝑝𝑞𝐼𝑦 − 𝑝𝑞𝐼𝑥 + 𝑟𝑞𝐼𝑥𝑧 )
The above equation can be simplified more, as shown below
1 2
𝑝̇ = 2
{𝐼𝑧𝑧 𝐿 + 𝐼𝑥𝑧 𝑁 − [𝐼𝑥𝑧 (𝐼𝑦𝑦 − 𝐼𝑥𝑥 − 𝐼𝑧𝑧 )𝑝 + (𝐼𝑥𝑧 + 𝐼𝑧𝑧 (𝐼𝑧𝑧 − 𝐼𝑦𝑦 )) 𝑟] 𝑞}
𝐼𝑥𝑥 𝐼𝑧𝑧 − 𝐼𝑥𝑧
1
𝑞̇ = [𝑀 + 𝐼𝑥𝑧 (𝑟 2 − 𝑝2 ) − 𝑝𝑟(𝐼𝑥𝑥 − 𝐼𝑧𝑧 )]
𝐼𝑦𝑦
1 2
𝑟̇ = 2
{𝐼𝑥𝑧 𝐿 + 𝐼𝑥𝑥 𝑁 + [𝐼𝑥𝑧 (𝐼𝑦𝑦 − 𝐼𝑥𝑥 − 𝐼𝑧𝑧 )𝑟 + (𝐼𝑥𝑧 + 𝐼𝑥𝑥 (𝐼𝑥𝑥 − 𝐼𝑦𝑦 )) 𝑝] 𝑞}
𝐼𝑥𝑥 𝐼𝑧𝑧 − 𝐼𝑥𝑧
If in addition we have 𝐼𝑥𝑧 = 0 then

𝑝̇ (𝑡) = [𝐿(𝑡) − 𝑞(𝑡)𝑟(𝑡) (𝐼𝑧 (𝑡) − 𝐼𝑦 (𝑡))] /𝐼𝑥 (𝑡)


𝑞̇ (𝑡) = [𝑀(𝑡) − 𝑝(𝑡)𝑟(𝑡)(𝐼𝑥 (𝑡) − 𝐼𝑧 (𝑡))]/𝐼𝑦 (𝑡)
𝑟̇ (𝑡) = [𝑁(𝑡) − 𝑝(𝑡)𝑞(𝑡) (𝐼𝑦 (𝑡) − 𝐼𝑥 (𝑡))] /𝐼𝑧 (𝑡)

• Now we want to express the angular velocity 𝛚 of the rotating frame in terms of
rotation angles (or Euler angles). Because the two reference frames (Earth-Body)
are rotating to each other, so the direction matrix 𝐑(𝜃, 𝜙, 𝜓) is a function of time.
cos 𝜃 cos 𝜓 cos 𝜃 sin 𝜓 − sin 𝜃
𝑹(𝜃, 𝜙, 𝜓) = [sin 𝜙 sin 𝜃 cos 𝜓 − cos 𝜙 sin 𝜓 sin 𝜙 sin 𝜃 sin 𝜓 + cos 𝜙 cos 𝜓 sin 𝜙 cos 𝜃 ]
cos 𝜙 sin 𝜃 cos 𝜓 + sin 𝜙 sin 𝜓 cos 𝜙 sin 𝜃 sin 𝜓 − sin 𝜙 cos 𝜓 cos 𝜙 cos 𝜃
The rotation matrix 𝑹 (or direction cosine matrix) is necessary to transform vectors
and point coordinates from the body fixed frame 𝐱 𝐵 to the navigation frame 𝐱 𝐸 and
vice versa: 𝐱 𝐵 = 𝑹(𝜃, 𝜙, 𝜓)𝐱 𝐸 for positive sense; or 𝐱 𝐸 = 𝑹(𝜃, 𝜙, 𝜓)𝐱 𝐵 for negative sense.
Let we take the derivative of 𝐫1 = 𝐑(𝜃, 𝜙, 𝜓)𝐫0 with respect to time. We get
𝑑𝐫1 𝑑 𝑑𝐑 𝑑𝐫0 𝑑𝐑 𝑑𝐑 𝑇
= [𝐑(𝜃, 𝜙, 𝜓)𝐫0 ] = 𝐫0 + 𝐑 = 𝐫0 = { 𝐑 } 𝐫1 with 𝐫0 = 𝐑−1 𝐫1 = 𝐑𝑇 𝐫1
𝑑𝑡 𝑑𝑡 𝑑𝑡 𝑑𝑡 𝑑𝑡 𝑑𝑡
Since 𝐫0 is independent on time (i.e. 𝐫0 in a body-fixed frame). From the other side
we know that (i.e. we consider the negative rotation −𝛚)
0 −𝜔𝑧 𝜔𝑦
𝑑𝐫1 𝑑𝐑 𝑇
= −𝛚 × 𝐫1 = −𝐖𝐫1 ⟹ 𝐖 = − 𝐑 = [ 𝜔𝑧 0 −𝜔𝑥 ]
𝑑𝑡 𝑑𝑡 −𝜔𝑦 𝜔𝑥 0
So by the chain rule we have
0 −𝜔𝑧 𝜔𝑦
𝑑𝐑 𝜕𝐑 𝑑𝜃 𝜕𝐑 𝑑𝜙 𝜕𝐑 𝑑𝜓 𝜕𝐑 𝑑𝜃 𝜕𝐑 𝑑𝜙 𝜕𝐑 𝑑𝜓 𝑇
= + + ⟹ [ 𝜔𝑧 0 −𝜔𝑥 ] = − { + + }𝐑
𝑑𝑡 𝜕𝜃 𝑑𝑡 𝜕𝜙 𝑑𝑡 𝜕𝜓 𝑑𝑡 −𝜔𝑦 𝜔𝑥 0 𝜕𝜃 𝑑𝑡 𝜕𝜙 𝑑𝑡 𝜕𝜓 𝑑𝑡

clear all,clc, syms th ph ps th1 ph1 ps1 real;


R1=[cos(ps) sin(ps) 0;-sin(ps) cos(ps) 0;0 0 1];
R2=[cos(th) 0 -sin(th);0 1 0;sin(th) 0 cos(th)];
R3=[1 0 0;0 cos(ph) sin(ph);0 -sin(ph) cos(ph)];
R=R3*R2*R1; R=simplify(R); RI=simplify(inv(R));
Zero1=simplify(R'-RI) % verification of 𝑹−1 = 𝑹𝑇
D1=diff(R,ps); D1=simplify(D1);
D2=diff(R,th); D2=simplify(D2);
D3=diff(R,ph); D3=simplify(D3);
Omega = -(D1*ps1 + D2*th1 + D3*ph1)*R'; Omega=simplify(Omega)
Omega1(1)=Omega(3,2), Omega1(2)=Omega(1,3), Omega1(3)=Omega(2,1)
The output of this code gives the angular velocity tensor 𝑾
0 −𝜔𝑧 𝜔𝑦 0 𝜃̇𝑠𝜙 − 𝜓̇ 𝑐𝜙 𝑐𝜃 𝜃̇𝑐𝜙 + 𝜓̇ 𝑠𝜙 𝑐𝜃
0 𝜓̇ 𝑐 ̇
𝜙 𝜃 − 𝜃 𝑠𝜙
𝑐
𝑾 = [ 𝜔𝑧 −𝜔𝑥 ] = [ 0 𝜓̇𝑠𝜃 − 𝜙̇ ]
−𝜔𝑦 𝜔𝑥 0 −𝜃̇𝑐𝜙 − 𝜓̇ 𝑠𝜙 𝑐𝜃 𝜙̇ − 𝜓̇𝑠𝜃 0
Hence we have
0 −𝜔𝑧 𝜔𝑦 𝜔𝑥 1 0 − sin(𝜃) 𝜙̇
𝑑𝐑 𝑇
𝑾=− 𝐑 = [ 𝜔𝑧 0 −𝜔𝑥 ] ⟺ [𝜔𝑦 ] = [0 cos(𝜙) sin(𝜙) cos(𝜃) ] [ 𝜃̇ ]
𝑑𝑡 −𝜔𝑦 𝜔𝑥 0 𝜔𝑧 0 − sin(𝜙) cos(𝜙) cos(𝜃) 𝜓̇
−1
𝜙̇ 1 0 − sin(𝜃) 𝜔𝑥 1 sin(𝜙) tan(𝜃) cos(𝜙) tan(𝜃) 𝜔𝑥
[ 𝜃̇ ] = [0 cos(𝜙) sin(𝜙) cos(𝜃) ] [𝜔𝑦 ] = [0 cos(𝜙) − sin(𝜙) ] [𝜔𝑦 ]
𝜓̇ 0 − sin(𝜙) cos(𝜙) cos(𝜃) 𝜔𝑧 0 sin(𝜙) / cos(𝜃) cos(𝜙) / cos(𝜃) 𝜔𝑧

The nonlinear dynamic model of the surface to air


missile is described by the next set of coupled differential equations.
𝑤
𝑢̇ (𝑡) = F𝑥 (𝑡)/𝑚(𝑡) − (𝑞(𝑡)𝑤(𝑡) − 𝑟(𝑡)𝑣(𝑡)), [m/s2 ] 𝛼(𝑡) = tan−1 ( )
𝑢
𝑣̇ (𝑡) = F𝑦 (𝑡)/𝑚(𝑡) − (𝑟(𝑡)𝑢(𝑡) − 𝑝(𝑡)𝑤(𝑡)), 2
[m/s ] 𝑣
𝛽(𝑡) = sin−1 ( )
[m/s2 ] 𝑉𝑀
𝑤̇ (𝑡) = F𝑧 (𝑡)/𝑚(𝑡) − (𝑝(𝑡)𝑣(𝑡) − 𝑞(𝑡)𝑢(𝑡)), −1 (cos
𝛼𝑡 = cos 𝛼 cos 𝛽)
𝑝̇ (𝑡) = [𝐿(𝑡) − 𝑞(𝑡)𝑟(𝑡) (𝐼𝑧 (𝑡) − 𝐼𝑦 (𝑡))] /𝐼𝑥 (𝑡), [rad/s2 ] 2
𝑉𝑀 = 𝑢 + 𝑣 + 𝑤 2
2 2

𝑞̇ (𝑡) = [𝑀(𝑡) − 𝑝(𝑡)𝑟(𝑡)(𝐼𝑥 (𝑡) − 𝐼𝑧 (𝑡))]/𝐼𝑦 (𝑡), [rad/s2 ]


𝑟̇ (𝑡) = [𝑁(𝑡) − 𝑝(𝑡)𝑞(𝑡) (𝐼 (𝑡) − 𝐼 (𝑡))] /𝐼 (𝑡), [rad/s2 ] Or alternatively
𝑦 𝑥 𝑧

ϕ̇(𝑡) = 𝑝(𝑡) + (𝑞(𝑡)sin(ϕ) + 𝑟(𝑡)cos(ϕ))tan(θ), [rad/s] (deg/s) 𝑤


𝛼(𝑡) = tan−1 ( )
θ̇(𝑡) = 𝑞(𝑡)cos(ϕ) − 𝑟(𝑡)sin(ϕ), [rad/s] (deg/s) 𝑢
𝑣
{ ψ̇(𝑡) = (𝑞(𝑡)sin(ϕ) + 𝑟(𝑡)cos(ϕ))/cos(θ), [rad/s] (deg/s) 𝛽(𝑡) = tan−1 ( )
𝑢

• Kinematic model can be obtained by using the 𝐯𝐸 = [𝑥̇ 𝑦̇ 𝑧̇ ]𝑇 = 𝐑−1 [𝑢 𝑣 𝑤]𝑇


𝑥̇ = [cos 𝜃 cos 𝜓]𝑢 + [sin 𝜙 sin 𝜃 cos 𝜓 − cos 𝜙 sin 𝜓]𝑣 + [cos 𝜙 sin 𝜃 cos 𝜓 + sin 𝜙 sin 𝜓]𝑤
𝑦̇ = [cos 𝜃 sin 𝜓]𝑢 + [sin 𝜙 sin 𝜃 sin 𝜓 + cos 𝜙 cos 𝜓]𝑣 + [cos 𝜙 sin 𝜃 sin 𝜓 − sin 𝜙 cos 𝜓]𝑤
𝑧̇ = −𝑢 sin(𝜃) + 𝑣 sin(𝜙) cos(𝜃) + 𝑤 cos(𝜙) cos(𝜃)
𝑇
𝐅(𝑡) = [F𝑥 (𝑡), F𝑦 (𝑡), F𝑧 (𝑡)] : is the external force includes the sum of aerodynamic,
pressure thrust, and gravitational forces (expressed in body coordinate system).

F𝑥 (𝑡) = F𝐴𝑥 + F𝑝𝑥 + Fg𝑥 , F𝑦 (𝑡) = F𝐴𝑦 + F𝑝𝑦 + Fg𝑦 , F𝑧 (𝑡) = F𝐴𝑧 + F𝑝𝑧 + Fg𝑧

𝐌(𝑡) = [𝐿(𝑡), 𝑀(𝑡), 𝑁(𝑡)]𝑇 [N. m]: is the total moment vector expressed in the body
fixed frame (roll, pitch, and yaw), L(𝑡) = L𝐴 + L𝑝 , M(𝑡) = M𝐴 + M𝑝 , N(𝑡) = N𝐴 + N𝑝 .

𝐈(𝑡) = diag[𝐼𝑥 (𝑡), 𝐼𝑦 (𝑡), 𝐼𝑧 (𝑡)]: is the inertia matrix (i.e. 𝐼𝑥 (𝑡), 𝐼𝑦 (𝑡), 𝐼𝑧 (𝑡) are the diagonal
elements of inertia matrix when products of inertia are zero), [kg. m2 ].

θ(𝑡), ϕ(𝑡), ψ(𝑡): Euler angle rotation in (pitch, roll, yaw angles), [rad] (deg)
ψ̇, θ̇, ϕ̇: rates of change of Euler angles in yaw, pitch, and roll, [rad/s] (deg/s)
𝛼(𝑡): angle of attack, [rad](deg), with 𝛼(𝑡) = tan−1(𝑤/𝑢)
𝛽(𝑡): angle of sideslip, [rad](deg) and 𝛼𝑡 (𝑡): total angle of attack, [rad](deg).
■ The value of missile mass 𝑚 is the value determined by table lookup as a
function of the current simulation time, or, as an alternative, 𝑚 can be calculated
within the simulation, by using the impulse momentum theorem
1 𝑡
Total linear momentum = (𝑚0 − 𝑚(𝑡)) × 𝐼𝑠𝑝 ⟹ 𝑚(𝑡) = 𝑚0 − ∫ F (𝑡)𝑑𝑡 , [kg]
𝐼𝑠𝑝 0 pref
𝑚(𝑡) = instantaneous mass of missile, [kg].
𝑚0 = missile mass of the missile at time zero (i.e. ,at the time of launch), [kg]
Fpref (𝑡) = reference thrust force (measured at sea level), [N]
𝐼𝑠𝑝 = specific impulse of propellant [N. s/kg]

Fpref (𝑡) = −𝐼𝑠𝑝 (𝑑𝑚/𝑑𝑡) = 𝐼𝑠𝑝 (𝑑𝑚𝑒 /𝑑𝑡) ⟺ 𝑚(𝑡 + ∆𝑡) = 𝑚(𝑡) − Fpref (𝑡)∆𝑡/𝐼𝑠𝑝

Note: In rocketry the specific impulse as the impulse per unit mass of propellant
used is simply the effective exhaust velocity: 𝐼𝑠𝑝 = 𝑢eq [m/s] where 𝐼𝑠𝑝 is the specific
impulse of propellant, as defined above, and measured in meters per second and
𝑢eq is the effective exhaust velocity measured in [m/s]. It is related to the thrust, or
forward force on the rocket by the equation: Fpref (𝑡) = 𝐼𝑠𝑝 (𝑑𝑚𝑒 /𝑑𝑡) where 𝑑𝑚𝑒 /𝑑𝑡 is
the mass flow rate, which is minus the time-rate of change of the vehicle's mass,
since fuel is being expelled. In some references authors use the definition of
specific impulse as being the 𝐼𝑠𝑝 = 𝑢eq /g [s].

■ The input thrust table is used to look up the thrust Fpref corresponding to the
reference atmospheric pressure as a function of time. The thrust is corrected for
the ambient atmospheric pressure P𝑎 by: F𝑝 (𝑡) = Fpref + (Pref − P𝑎 )A𝑒 , [N]

or F𝑝 (𝑡) = 𝐼𝑠𝑝 𝑚̇𝑒 + (Pref − P𝑎 )A𝑒 = 𝑚̇𝑒 v𝑟𝑒 + (Pref − P𝑎 )A𝑒 , [N]

F𝑝 (𝑡) = instantaneous input thrust produced by missile engine, [N]


P𝑎 = ambient atmospheric pressure and Pref (𝑡) = reference ambient pressure.

■ Assuming that the moment of inertia varies linearly with the mass, the moment
of inertia is updated using:

𝐼(𝑡) = 𝐼0 − (𝐼0 − 𝐼𝑏0 )([𝑚0 − 𝑚(𝑡)]/[𝑚0 − 𝑚𝑏0 ]), [kg. m2 ]

𝐼0 = diag([𝐼𝑥0 𝐼𝑦0 𝐼𝑧0 ]) = moment of inertia at launch, [kg. m2 ]


𝐼𝑏0 = diag([𝐼𝑥𝑏0 𝐼𝑦𝑏0 𝐼𝑧𝑏0 ]) = moment of inertia at burnout, [kg. m2 ]
𝑚𝑏0 = missile mass at burnout, [kg]
𝐼(𝑡) = diag([𝐼𝑥 (𝑡) 𝐼𝑦 (𝑡) 𝐼𝑧 (𝑡)]) = instantaneous moment of inertia, [kg. m2 ]
■ The location of the center of mass varies linearly with the mass. The location of
the center of mass 𝑥𝑐𝑚 is updated by using
𝑥𝑐𝑚 (𝑡) = 𝑥𝑐𝑚0 − [𝑥𝑐𝑚0 − 𝑥𝑐𝑚𝑏0 ]([𝑚0 − 𝑚(𝑡)]/[𝑚0 − 𝑚𝑏0 ]), [m]
𝑥𝑐𝑚 (𝑡) = instantaneous distance from missile noze to center of mass, [m]
𝑥𝑐𝑚0 = distance from missile noze to center of mass at launch, [m]
𝑥𝑐𝑚𝑏0 = distance from missile noze to center of mass at burnout, [m]
𝑡 = simulated time, [s], and 0 < 𝑡 < 𝑡𝑏 (i.e. 𝑡𝑏 = the burnout time)

■ In most missiles the magnitude of the thrust 𝐅𝑃 , is directed along the 𝑥𝑏 -axis (the
missile centerline). Then, the components of the thrust vector 𝐅𝑃 expressed in the
body reference frame are:
𝐹𝑝𝑥 𝐹pref + (𝑃ref − 𝑃𝑎 )𝐴𝑒
𝐹
𝐅𝑃 = [ 𝑝𝑦 ] = [ 0 ], [N]
𝐹𝑝𝑧 0
𝐵
■ The components of the thrust moment vector 𝐌𝑃 are given by

𝐿𝑝 = 0, [N. m], 𝑀𝑝 = 𝐹𝑃𝑧 𝑙𝑝 , [N. m], 𝑁𝑝 = −𝐹𝑃𝑦 𝑙𝑝 , [N. m]


The distance 𝑙𝑝 is calculated from missile nose to current value center of mass 𝑥𝑐𝑚 .
■ The gravitational force expressed in body coordinates is calculated by multiplying
the vector 𝐅𝐠 by the matrix that transforms from earth frame to the body frame:
Fg𝑥 𝑏 Cθ Cψ Cθ Sψ −Sθ 0 −𝑚(𝑡)g sin θ
F
[ g𝑦 𝑏 ] = [Sϕ Sθ Cψ − Cϕ Sψ Sϕ Sθ Sψ + Cϕ Cψ Sϕ Cθ ] [ 0 ] = [𝑚(𝑡)g sin ϕ cos θ] , [N]
Fg𝑧 Cϕ Sθ Cψ + Sϕ Sψ Cϕ Sθ Sψ − Sϕ Cψ Cϕ Cθ 𝑚(𝑡)g 𝑚(𝑡)g cos ϕ cos θ
𝑏

■ The resultant (total) aerodynamic force 𝐅𝐴 on the missile can be resolved in any
coordinate frame to give three orthogonal components.
−𝐴 𝐹𝐿 sin 𝛼𝑡 − 𝐹𝐷 cos 𝛼𝑡
𝐹𝐴𝑥
√ 2 2
𝐅𝐴 = [𝐹𝐴𝑦 ] = [ −𝑁 (𝑣/ 𝑣 + 𝑢 ) ] = [ −𝑣(𝐹𝐷 sin 𝛼𝑡 + 𝐹𝐿 cos 𝛼𝑡 )/√𝑣 2 + 𝑢2 ] , [N]
𝐹𝐴𝑧 𝐵 −𝑁 (𝑤/√𝑣 2 + 𝑢2 ) −𝑤(𝐹𝐷 sin 𝛼𝑡 + 𝐹𝐿 cos 𝛼𝑡 )/√𝑣 2 + 𝑢2

Since the drag force 𝐅𝐷 is by definition directed opposite the velocity vector 𝐯𝑀 and
lift 𝐅𝐿 is by definition perpendicular to the velocity vector and lies in the plane
formed by the velocity vector and the normal force vector, the aerodynamic force is
given by
𝐅𝐴 = 𝐹𝐿 [(𝐮𝑉𝑀 × 𝑵) × 𝐮𝑉𝑀 ] − 𝐹𝐷 𝐮𝑉𝑀 , [N] with 𝐮𝑉𝑀 = 𝐯𝑀 /‖𝐯𝑀 ‖, [dimensionless]

Note: In the control process the calculated maneuver acceleration vector 𝐚c is


perpendicular to the tangent to the missile flight path (missile velocity vector)
rather than to the missile centerline. Since the component of aerodynamic force
perpendicular to the velocity vector is lift instead of normal force, certain equations
must be redefined for these simulations: 𝐅𝐿 = 𝑚𝐚c , [N], this gives the lift force
vector that is required to produce the commanded-lateral-acceleration vector 𝐚c . It
is assumed that the missile responds as necessary (within any stated limits) to
achieve this lift force. (i.e. the needed 𝐚c is produced by the autopilot)
Equations for calculating the aerodynamic forces and moments affecting the
missile are given by:

Drag: 𝐹𝐷 = 0.5𝜌𝑉 2 𝐶𝐷 S, [N] Lift: 𝐹𝐿 = 0.5𝜌𝑉 2 𝐶𝐿 S, [N] Side Force: 𝐹𝑆 = 0.5𝜌𝑉 2 𝐶𝑌 S, [N]
L𝐴 = 0.5𝜌𝑉 2 𝐶𝑙 Sd, [N. m] M𝐴 = 0.5𝜌𝑉 2 𝐶𝑚 Sd, [N. m] N𝐴 = 0.5𝜌𝑉 2 𝐶𝑛 Sd, [N. m]

𝐹𝐷 magnitude of aerodynamic drag force vector 𝐅𝐷 , [N]


𝐹𝐿 magnitude of aerodynamic lift force vector 𝐅𝐿 , [N]
𝐿𝐴 , 𝑀𝐴 , 𝑁𝐴 : components of aerodynamic moment 𝐌𝑨 , in body coordinate frame, [N. m]
𝐶𝐷 , 𝐶𝐿 : aerodynamic drag, lift, coefficient, [dimensionless]
𝐶𝑙 : aerodynamic roll moment coefficient about center of mass, [dimensionless]
𝐶𝑚 : aerodynamic pitch moment coefficient about center of mass, [dimensionless]
𝐶𝑛 : aerodynamic yaw moment coefficient about center of mass, [dimensionless]
𝑆: aerodynamic reference area, [m2 ]
𝜌: Standard atmospheric density at sea level, [kg/m3 ].
𝑑: aerodynamic reference length of body, [m]
𝑉: speed of body, speed of air relative to body, magnitude of velocity vector 𝐯⃗⃗, [m/s]

■ The aerodynamic coefficients are given by:


𝐶𝐷 = 𝐶𝐷0 + 𝑘𝐶𝐿 2 , [dimensionless]
𝐶𝐿 = 𝐶𝐿𝛼 𝛼, [dimensionless]
𝑑
𝐶𝑙 = 𝐶𝑙𝛿 𝛿𝑟 + (𝐶 𝑝), [dimensionless]
2V 𝑙𝑝
𝑥𝑐𝑚 − 𝑥ref 𝑑
𝐶𝑚 = 𝐶𝑚 ref − 𝐶𝑁 𝑧 ( )+ (𝐶 + 𝐶𝑚𝛼̇ ) 𝑞, [dimensionless]
𝑑 2V 𝑚𝑞
𝑥𝑐𝑚 − 𝑥ref 𝑑
𝐶𝑛 = 𝐶𝑛 ref + 𝐶𝑁 𝑦 ( )+ (𝐶 + 𝐶𝑛𝛽̇ ) 𝑟, [dimensionless]
𝑑 2V 𝑛𝑟
𝐶𝐷0 : zero-lift drag coefficient [dimensionless]
𝐶𝐿𝛼 : slope of curve formed by lift coefficient 𝐶𝐿 versus 𝛼, [rad−1 ](deg −1 )
𝐶𝑙𝑝 : roll damping derivative, [rad−1 ](deg −1 )
𝐶𝑙𝛿 : slope of curve formed by roll moment coefficient 𝐶𝑙 versus 𝛿𝑟 . [rad−1 ](deg −1 )
𝑘: constant depending on body shape and flow regime, [dimensionless]
δ𝑟 , δ𝑝 , δ𝑦 angles of effective control surface deflections in (roll, pich, yaw ) directions [rad] (deg)
𝐶𝑚 ref : pitching moment coefficient about reference moment station, [dimensionless]
𝐶𝑛 ref : yawing moment coefficient about reference moment station, [dimensionless]
𝐶𝑁 𝑦 : coefficient corresponding to the normal force on 𝑦𝑏 -axis, [dimensionless]
𝐶𝑁 𝑧 : coefficient corresponding to the normal force on 𝑧𝑏 -axis, [dimensionless]
𝐶𝑛 𝑟 : yaw damping derivative relative to yaw rate 𝑟, [rad−1 ](deg −1 )
𝐶𝑛 ̇ : yaw damping derivative relative to angle-of sideslip rate 𝛽, [rad−1 ](deg −1 )
𝛽

𝐶𝑚𝑞 : pitch damping derivative relative to pitch rate 𝑞, [rad−1 ](deg −1 )


𝐶𝑚𝛼̇ : pitch damping derivative relative to angle of attack rate 𝛼, [rad−1 ](deg −1 )
𝑥𝑐𝑚 : instantaneous distance from missile nose to center of mass, [m]
𝑥ref : distance from missile nose to reference moment station, [m].
The pitch and yaw components of the aerodynamic moment about the moment
reference station are calculated by using
𝐶𝑚 ref = 𝐶𝑚𝛼 𝛼 + 𝐶𝑚𝛿 𝛿𝑝 , [dimensionless] 𝐶𝑛 ref = 𝐶𝑛𝛽 𝛽 + 𝐶𝑛𝛿 𝛿𝑦 , [dimensionless]

𝐶𝑛𝛽 : slope of curve formed by yawing moment coefficient 𝐶𝑛 versus 𝛽, [rad−1 ](deg −1 )
𝐶𝑛𝛿 : slope of curve formed by yawing moment coefficient 𝐶𝑛 versus δ𝑦 . [rad−1 ](deg −1 )
𝐶𝑚𝛼 : slope of curve formed by pitch moment coefficient 𝐶𝑚 versus 𝛼, [rad−1 ](deg −1 )
𝐶𝑚𝛿 : slope of curve formed by pitch moment coefficient 𝐶𝑚 versus δ𝑝 . [rad−1 ](deg −1 )

Since the missile is assumed to have cruciform symmetry, therefore


𝐶𝑛𝛽 = 𝐶𝑚𝛼 , [rad−1 ](deg −1 ), 𝐶𝑛𝑟 = 𝐶𝑚𝑞 , [rad−1 ](deg −1 )
𝐶𝑛𝛿 = 𝐶𝑚𝛿 , [rad−1 ](deg −1 ), 𝐶𝑛𝛽̇ = 𝐶𝑚𝛼̇ , [rad−1 ](deg −1 )

Coefficients corresponding to the components of the normal force in the 𝑦𝑏 - and 𝑧𝑏 -


axes are calculated by 𝐶𝑁 𝑦 = F𝐴𝑦 /𝑄𝑆, [dimensionless]. 𝐶𝑁 𝑧 = F𝐴𝑧 𝑏 /𝑄𝑆, [dimensionless]
𝑏
with 𝑄: dynamic pressure parameter, [Pa] and 𝑆: aerodynamic reference area, [m2 ].
The term 𝑄 = 0.5𝜌𝑉 2 is a very important quantity, known as the dynamic pressure
parameter which is equal to the kinetic energy per unit volume of air. Two
equivalent forms of 𝑄 will be presented here: 𝑄 = 0.5𝜌𝑉 2 = 0.7𝜌𝑃𝑎 𝑀𝑁2 where
𝑀𝑁 =Mach number, [dimensionless]. 𝑃𝑎 =atmospheric pressure ,[Pa]

The value of the aerodynamic force coefficients for a given body configuration is
affected primarily by the shape of the body (including any control-surface
deflections), the orientation of the body within the flow (angle of attack), and the
flow conditions. The flow conditions can be specified by two parameters: Mach
number and Reynolds number. As the missile speed approaches and exceeds the
speed of sound, the compressibility characteristics of the air have a pronounced
effect on the aerodynamic forces and moments, 𝑀𝑁 = 𝑉𝑀 /𝑉𝑆 . Aerodynamic flow over
a body has different characteristics, depending on the speed of the air relative to
the body. The aerodynamic coefficients in turn depend on these characteristics of
the flow. The different flow characteristics are grouped into five basic flow regimes
based on Mach number 𝑀𝑁 , these regimes are described as:
1. Incompressible subsonic flow: 0.0 < 𝑀𝑁 < 0.3,
2. Compressible subsonic flow: 0.5 ≤ 𝑀𝑁 < 0.8,
3. Transonic flow: 0.8 ≤ 𝑀𝑁 < 1.2,
4. Supersonic flow: 1.2 ≤ 𝑀𝑁 < 5.0,
5. Hypersonic flow: 5.0 ≤ 𝑀𝑁 .
The Reynolds number is a measure of the ratio of the inertial properties of the fluid
flow to the viscous properties. The Reynolds number is given by: 𝑅𝑒 = 𝜌𝑉𝑑/𝜇
where: 𝑑 is the aerodynamic reference length of body. 𝑉 is the speed of a body. 𝜇 is
the atmospheric dynamic viscosity. 𝜌 is the atmospheric density. The reference
length 𝑑 is a scale factor that accounts for the effect of the size of the missile on
the flow characteristics. The missile diameter or its length is often selected as the
reference length. Force coefficients are functions of Reynolds number.
■ (The atmospheric data) In applications in which measured atmospheric data are
available at only one or a few altitudes, the atmosphere is modeled in a flight
simulation by using equations that extrapolate or interpolate data according to
known principles of atmospheric variation with altitude. These equations are

𝑇 = 𝑇1 + 𝑎(ℎ − ℎ1 ), [K]. 𝑃 = 𝑃1 + (𝑇/𝑇1 )−g0/(𝑎𝑅) , [Pa]


𝜌 = 𝑃/(𝑅𝑇), [kg/m3 ]. 𝑉𝑠 = √𝛾𝑅𝑇, [m/s]

a: lapse rate, [K/m] and 𝜌: atmospheric density, [kg/m3 ].


g 0 : magnitude of the vector 𝐠 ⃗⃗ 𝟎 due to gravity at the earth surface, [m/s 2 ]
ℎ: altitude at which atmospheric properties are to be calculated above sea level, [m]
ℎ1 : reference altitude (e.g., sea level or earth surface), [m]
𝑃: pressure at altitude ℎ, [Pa] and 𝑃1 : given pressure at altitude ℎ1 , [Pa]
𝑅: gas constant (287.05), [N. m]/(kg. K)
𝑇: temperature at altitude ℎ, [K] and 𝑇1 : given temperature at altitude ℎ1 , [K]
𝑉𝑠 :speed of sound at altitude ℎ, [m/s] and 𝛾: ratio of specific heat, [dimensionless]

In general, the equations of motion that describe the flight


of a missile are equally applicable to an airborne target; however, the equations
used to calculate target motion in a missile flight simulation are usually greatly
simplified. If the objective of the simulation is to calculate the maximum defended
area covered by a particular surface-to-air missile, straight, constant-speed target
flight paths may be sufficient. At the other extreme, however, if the performance of
the missile is to be studied when it engages a particular type of aircraft as it
performs specified evasive maneuvers, a much more detailed model of the target is
required. When the simulated target is to fly a straight, constant speed flight path,
the target position vector at time 𝑡 is calculated by using

𝐩𝑇 = 𝐩𝑇0 + 𝐯𝑇 𝑡 [m]

𝐩𝑇 : is the position vector of the target [m], 𝐩𝑇0 is initial position vector of the target.
𝐯𝑇 : is the velocity vector (assumed constant) of target center of mass [m/s].

When an aircraft is flown through defended airspace, the pilot may perform evasive
maneuvers to make it more difficult for defensive gunfire or missiles to intercept
his aircraft. If the pilot is aware that he is being engaged by a particular type of
missile, he may perform evasive maneuvers prescribed for use against that
particular type of missile. To be most effective, the timing and direction-or
directions of compound maneuvers-may be important. The magnitudes of the
accelerations of evasive maneuvers are particularly important. When a pilot is not
aware of a specific engagement by defensive fire, he may perform a more or less
continuous series of maneuvers, called jinking, while flying through known
defended regions. Other examples of target maneuvers that might be included in a
missile flight simulation are terrain-following and terrain-avoidance flight paths or
map-of-the-earth fright paths flown by helicopters for concealment.
Usually, the fidelity required to model the target flight path is insufficient to
warrant the use of sophisticated numerical integration techniques for solving the
equations of motion. The improved Euler method is commonly used to update
target position and velocity from one calculation time to the next. By employing
this” method, the target position is updated by using:
𝐩𝑇 (𝑡) = 𝐩 𝑇 (𝑡 − Δ𝑡) + 𝐯𝑇 (𝑡 − Δ𝑡)Δ𝑡 + 𝐚 𝑇 (Δ𝑡 2 /2) [m]
𝐚𝑇 = total acceleration vector of target, [m/s2 ]
𝐩𝑇 = position vector of target, [m] and Δ𝑡 = computation time step, [s]
𝐯𝑇 = velocity vector (assumed constant) of target center of mass, [m/s].
(𝑡) = indicate that the associated variable is calculated at the current time, [s]
(𝑡 − Δ𝑡) = indicate that the associated variable was calculated at the previous time,

The velocity vector 𝐯𝑇 (𝑡) at the end of the current calculation interval is given by
the equation: 𝐯𝑇 (𝑡) = 𝐯𝑇 (𝑡 − Δ𝑡) + 𝐚 𝑇 Δ𝑡, [m/s]. The target acceleration vector 𝐚 𝑇 is
calculated by using: 𝐚𝑇 = 𝛚𝑇 × 𝐯𝑇 (𝑡 − Δ𝑡), [m/s 2 ]
Where 𝛚𝑇 = the angular rate vector of the target
flight path, [deg/s] (rad/s) . The method of
calculating the flight path angular rate vector varies
depending on the type of target maneuver, The
fright path angular rate vector for controlling target
maneuvers can be input as a constant or as a
tabular function of time, or it can be calculated
within the simulation. The magnitudes of target
maneuvers are usually specified in terms of
something called the load factor 𝑛g = 𝐿/𝑊, where
𝑊 = 𝑚g is the weight of the aircraft. In a horizontal turn, bank angle 𝜙𝑇 is a
function of only the load factor. Given the load factor, the bank angle can be
calculated by using: 𝜙𝑇 = cos−1 (1/𝑛g ), [rad] Where: 𝜙𝑇 =target aircraft bank angle
(Euler roll angle of the target coordinate system relative to the earth system),
[rad] (deg). For horizontal, constant-speed, coordinated turns, the magnitude of the
total acceleration of the aircraft can be calculated for any given load factor by
1/2
using aT = g[𝑛g2 − 1] , [m/s2 ] where aT is the magnitude of 𝐚⃗⃗𝑇 of the target. For
horizontal turns the angular rate vector is given by: 𝝎T = [0 0 aTc /vT ], [rad/s]
where: aTc = commanded target maneuver acceleration, [m/s2 ]. vT = magnitude of
target velocity vector (i.e. constant), [m/s]. A positive value of 𝐚𝑇𝐶 produces right-
hand turns; a negative value produces left-hand turns. If changes in speed are
desired during the turn, the magnitude of target velocity 𝐯𝑇 is varied accordingly.
Although pilots employ different types of jinking maneuvers, one that is commonly
employed in simulations is a simple weaving flight path in a horizontal plane. The
weaving flight path can be modeled as a cosine curve by calculating the maneuver
acceleration: a 𝑇𝐶 = amax cos((2𝜋/𝑃𝑑 )𝑡𝑚𝑖 ) [m/s 2 ] where amax is the magnitude of max-
maneuver acceleration. a 𝑇𝐶 is the commanded target maneuver acceleration. 𝑃𝑑 is
the period of target wave maneuver [s] and 𝑡𝑚𝑖 is time since starting maneuver [s].
If horizontal, coordinated maneuvers are assumed, the maximum maneuver
2 1/2
acceleration is calculated by using amax = g[𝑛gmax − 1] . The time 𝑡𝑚𝑖 is calculated
as the difference between current simulation time 𝑡 and the time when the
maneuver was initiated. The maneuver initiation time may be input, or it can be
calculated within the simulation as a function of the engagement, such as the time
when the missile reaches a specified range from the target. To avoid the
discontinuities in the commanded target maneuver acceleration we pass a 𝑇𝐶
through a first-order transfer function (low-pass filter) before using it in a digital
simulation. The following steps summarize the curved path for target maneuver
aTc = amax cos((2𝜋/𝑃𝑑 )𝑡𝑚𝑖 ) [m/s 2 ]
a 𝑇𝑎𝑐ℎ (𝑡) = a 𝑇𝑎𝑐ℎ (𝑡 − ∆𝑡) exp(∆𝑡/𝜏) + aTc [1 − exp(∆𝑡/𝜏)] [m/s 2 ]
𝝎T = [0 0 a 𝑇𝑎𝑐ℎ /vT ], [rad/s]
𝐚𝑇 = 𝛚𝑇 × 𝐯𝑇 (𝑡 − Δ𝑡), [m/s2 ]
𝐯𝑇 (𝑡) = 𝐯𝑇 (𝑡 − Δ𝑡) + 𝐚𝑇 Δ𝑡, [m/s]
𝐚 𝑇 Δ𝑡 2
𝐩𝑇 (𝑡) = 𝐩𝑇 (𝑡 − Δ𝑡) + 𝐯𝑇 (𝑡 − Δ𝑡)Δ𝑡 + [m]
2
■ (The missile position) The differential equations of missile velocity yielding the
components of the missile position vector 𝐯𝑀 (𝑡) = 𝐩̇ 𝑀 (𝑡) in the earth coordinate
system 𝐩𝑀 (𝑡) = 𝐩𝑀 (𝑡 − Δ𝑡) + 𝐯𝑀 (𝑡 − Δ𝑡)Δ𝑡 where
v𝑀𝑥 Cθ Cψ Cθ Sψ −Sθ 𝑢
(𝐯𝑀 (𝑡))𝐸 = v𝑀𝑥 𝐢 + v𝑀𝑦 𝐣 + v𝑀𝑧 𝐤 = [v𝑀𝑦 ] = [Sϕ Sθ Cψ − Cϕ Sψ Sϕ Sθ Sψ + Cϕ Cψ Sϕ Cθ ] [ 𝑣 ]
v𝑀𝑧 𝐸 Cϕ Sθ Cψ + Sϕ Sψ Cϕ Sθ Sψ − Sϕ Cψ Cϕ Cθ 𝑤 𝐵
The equation (𝐯𝑀 )𝐸 = 𝐑(𝐯𝑀 )𝐵 is called kinematic model (i.e. navigation equations).

■ (The relative motion) The relative range vector 𝐫 is the vector extending from the
position of the missile center of mass to the position of the target center of mass.
The relative range vector is identical to the line-of-sight vector from the missile to
the target if the distances from the missile seeker to the missile center of mass and
the distance from the track point to the target center of mass are neglected.
Calculations of seeker tracking and of miss distance depend on the relative range
vector. The position of the target relative to the missile is defined by the relative
range vector 𝐫, which is calculated in the earth frame by using: 𝐫 = 𝐩𝑇 − 𝐩𝑀 , [m]
where 𝐩𝑀 is the position vector of the missile, [m] and 𝐩𝑇 is the position vector of
the target, [m] and 𝐫 is the range vector from missile to target, [m].

In case of perfect seeker track, the angular rate of the seeker is set equal to the
angular rate of 𝐫, and the angle between 𝐫 and the missile centerline axis is the
seeker gimbal angle. The relative velocity vector is the difference between the target
and missile velocity vectors, i.e. 𝐯𝑇/𝑀 = 𝐯𝑇 − 𝐯𝑀 , [m] where 𝐯𝑀 is the velocity vector
of the missile [m/s], and 𝐯𝑇 is the velocity vector of the target, [m/s] and 𝐯𝑇/𝑀 is
the velocity vector of target center of mass relative to missile center of mass,[m/s].
Normalizing the relative velocity vector 𝐯𝑇/𝑀 gives the unit vector, 𝐮𝑇/𝑀 that is
𝐮𝑇/𝑀 = 𝐯𝑇/𝑀 /‖𝐯𝑇/𝑀 ‖ = (𝐯𝑇 − 𝐯𝑀 )/‖𝐯𝑇 − 𝐯𝑀 ‖ which has the direction of 𝐯𝑇/𝑀 .
■ (Miss distance) Miss distance 𝐌𝑑 = 𝐫 − (𝐫. 𝐮𝑇/𝑀 )𝐮𝑇/𝑀 is often used as a measure of
missile system performance. In general, the smaller the miss distance, the greater
the probability of killing the target. Missile kill probability is a function of many
factors-including 𝐌𝑑 , fuze, warhead characteristics, and target vulnerability.
Acceptable miss distance (sufficiently small) is the first criterion of a successful
engagement because the fuze and warhead must be delivered relatively close to the
target in order to perform their functions. Miss distance is usually defined as the
closest approach of some point on the missile-usually the missile center of mass-to
some point on the target-often the target center of mass. When miss distance is
defined as the closest approach of the missile center of mass to the target center of
mass, the closest approach occurs when the range vector 𝐫 reaches a minimum.

The autopilot is a link between the function that


indicates a change of heading is needed (guidance processor) and the mechanism
that can change the heading (control system). The "autopilot" receives guidance
commands and processes them to the controls such as deflections or rates of
deflection of control surfaces or jet controls. The control subsystem transfers the
autopilot commands to aerodynamic or jet control forces and moments to change
the position of the airframe, to attain the commanded maneuver by rotating the
body of a missile to a desired angle of attack. The design of the autopilot depends
on the aerodynamics of the missile airframe and the type of controls employed. The
autopilot may contain amplifiers, integrators, and mixing circuits that send signals
to the proper control surface actuators.
flight path commanded
Guidance processor and error acceleration Control
(part of seeker) Autopilot system

The autopilot in a missile has two basic


functions: to ensure stable flight and to translate the guidance law into control-
surface deflection commands. It is assumed that the studied missile has four
control surfaces in a cruciform pattern and that commanded control-surface
deflections are proportional to commanded lateral accelerations for maneuver
commands and proportional to commanded roll rates for roll
commands. In six degree of freedom model, the pitch and yaw 𝛿4 𝛿1
channels can be considered as decoupled SISO systems; this
simplifies their design. A constant roll position creates normal
working conditions for missile components. The relationship × 𝑦𝑏
between the actuator commands 𝛿𝑃 , 𝛿𝑌 , 𝛿𝑅 and the individual
fin deflections is given for this model as follow 𝛿𝑃 = (𝛿2 − 𝛿4 )/2,
𝑧𝑏 𝛿2
𝛿𝑌 = (𝛿1 − 𝛿3 )/2, and 𝛿𝑅 = (𝛿1 + 𝛿2 + 𝛿3 + 𝛿4 )/4 where 𝛿𝑃 is the 𝛿3
autopilot pitch fin command, 𝛿𝑌 is the autopilot yaw fin command, 𝛿𝑅 is the
autopilot roll fin commands. 𝛿𝑖 is the deflection angle of the 𝑖 𝑡ℎ control surface,
i=1,2,3,4. In five degree of freedom model, the pitch and yaw channels are coupled
and roll rate is set to zero, thus the actuator commands are 𝛿𝑃 = 𝛿2 , 𝛿𝑌 = 𝛿1 , 𝛿𝑅 = 0.
The restrictions related to the fin deflections are transformed into the autopilot
limits (𝛿𝑃 , 𝛿𝑌 , 𝛿𝑅 ), which, in turn, impose constrains on the missile acceleration.
Once the guidance processor has determined the magnitude and direction of the
error in the missile flight path and the autopilot has determined the steering
command, the missile control system must adjust the control surfaces to produce
the acceleration required to correct the flight path. This corrective acceleration is
applied in a lateral direction (perpendicular to the missile flight path) to change the
direction of the missile velocity vector. A moment, is required to cause a missile to
rotate by angle of attack, this moment can be developed by means of fin deflection.

The designs of the control system actuators of different missiles may vary
considerably, but all have a common purpose, i.e., to convert autopilot commands
into fin deflections. The input to the model is the control--surface deflection
command; the output is the control-surface defection achieved. The commanded-
lateral-acceleration vector 𝐚𝑐 is an essential design parameter and is computed by
one type of the known guidance control laws such as the line of sight (beam rider)
control or proportional navigation control. The vector 𝐚𝑐 is transformed to the body
reference frame by 𝐚𝑐𝑏 = [𝑻𝑒/𝑏 ]𝐚𝑐 = [𝐴𝑐1 𝐴𝑐2 𝐴𝑐3 ]𝑇 . The actuator piston pressure
for control surfaces 1 and 2 are given by 𝑃act1 = −𝐺𝑝 𝐴𝑐3 , 𝑃act2 = 𝐺𝑝 𝐴𝑐2 where 𝑃act1 is
the actuator pressure for the pitch channel [pa] and 𝑃act2 is the actuator pressure
for the yaw channel [pa]. 𝐺𝑝 is the gain factor relating the actuator pressure to
acceleration command [pa. s2 /m]. The control-surface angles of attack that result
from the actuator pressures commanded by autopilot are given by 𝛼𝑝 = −𝐺𝑛 𝐴𝑐3 /𝑄,
and 𝛼𝑦 = 𝐺𝑛 𝐴𝑐2 /𝑄 where 𝐺𝑛 is a gain factor relating the angle of attack to
acceleration command per unit dynamic pressure [rad. pa. s2 /m]. 𝛼𝑝 commanded
angle of attack of pitch-channel control surface [rad]. 𝛼𝑦 commanded angle of
attack of yaw-channel control surface [rad].

The seeker tracking point is assumed to be


the center of mass of the target and although a first-order lag in the tracking rate
is introduced later, the small angular deviation of the seeker boresight-axis vector
from the line of sight (LoS) to the tracking point is not calculated. Also the
displacement of the physical position of the seeker from the missile center of mass
is considered negligible for this application. Therefore, the seeker LoS vector 𝛔 is
assumed to be identical with the range vector: 𝛔 ≈ 𝐫 [m]. The angular rate of the
seeker LoS vector is calculated by using 𝝎𝝈 = (𝛔 × 𝐯𝑇/𝑀 )/‖𝛔‖2 ≈ 𝝎gl [rad/s]. The
angular rate of the seeker head lags the angular rate of the line-of-sight vector.
This lag is taken into account in calculating the guidance commands. This lag is
assumed to be represented by a first-order transfer function, and the seeker-head
rate is calculated by 𝝎ach (𝑡) = 𝝎ach (𝑡 − ∆𝑡) exp(∆𝑡/𝜏1 ) + 𝝎𝝈 (1 − exp(∆𝑡/𝜏1 )) [rad/s].
There are assumed to be delays involved in processing the seeker-head angular
rate signal. The filtered seeker-head angular rate signal is given by the following
relation 𝝎f (𝑡) = 𝝎f (𝑡 − ∆𝑡) exp(∆𝑡/𝜏2 ) + 𝝎ach (1 − exp(∆𝑡/𝜏2 )) [rad/s] where 𝜏1 : seeker
tracking loop constant. 𝜏2 : seeker signal processing time constant.
The dynamic response of the autopilot-control system is assumed to be described
by a first-order system with time constant 𝜏3 . The achieved angles of attack of the
fins will lag those given in 𝛼𝑝 and 𝛼𝑦 as given by

𝛼𝑝𝑎 (𝑡) = 𝛼𝑝𝑎 (𝑡 − ∆𝑡) exp(−∆𝑡/𝜏3 ) + 𝛼𝑝 [1 − exp(−∆𝑡/𝜏3 )] [rad]


𝛼𝑦𝑎 (𝑡) = 𝛼𝑦𝑎 (𝑡 − ∆𝑡) exp(−∆𝑡/𝜏3 ) + 𝛼𝑦 [1 − exp(−∆𝑡/𝜏3 )] [rad]

𝛼𝑝𝑎 is the achieved angle of attack of pitch-channel control surface [rad]. 𝛼𝑦𝑎 is the
achieved angle of attack of yaw-channel control surface [rad].

The angles of incidence of the control fins, i.e., the fin deflection angles relative to
the missile body, depend on the fin angles of attack and on the missile body angles
of attack and sideslip. The fin angles of incidence are calculated by using
equations 𝛿𝑝 = 𝛼𝑝𝑎 − 𝛼 and 𝛿𝑦 = 𝛼𝑦𝑎 − 𝛽. The absolute values of 𝛿𝑝 and 𝛿𝑦 are tested
against the maximum fin deflection angle 𝛿𝑚𝑎𝑥 if the absolute value of a fin angle of
incidence exceeds 𝛿𝑚𝑎𝑥 it is reset to 𝛿𝑚𝑎𝑥 and retains its original sign.

The designs of the control system components-power


sources, power transmission media, servos, and actuators of different missiles may
vary considerably, but all have a common purpose,
i.e., to convert autopilot commands into fin 𝛿 𝑐 error 𝐾 𝛿
deflections. The input to the closed loop servo- 𝑠
system is the control-surface deflection command;
the output is the control-surface defection achieved. Feedback Loop
The relationship between the output and the input is
defined mathematically by appropriate transfer functions and logical elements
such as limits on the magnitudes of control-surface defections. The transfer
function of the unity feedback control system (i.e. a system consisting of only the
servo-motor with no feedback) is 𝐺(𝑠) = 𝛿(𝑠)/𝑒(𝑠) where 𝑠 is the Laplace variable,
𝛿(𝑠) is the achieved control-surface deflection, [rad(deg)]. 𝑒(𝑠) = 𝛿𝑐 (𝑠) − 𝛿(𝑠) is the
error deflection, and 𝐺(𝑠) = 𝐾/𝑠 is the open loop control system transfer function
(i.e. the servo-model). Generally, the servomotors are modeled by first order
transfer function given by 𝐺(𝑠) = 𝐾/𝑠. The closed loop transfer function is given by
𝐺𝑐 (𝑠) = 1/(1 + 𝜏𝑠) where 𝜏 = 1/𝐾, is the control system time constant depends on
the type of motors and 𝐾 is the proportionality constant.

Target flight Missile


Intercept Heading
Position
error 𝑒 Guidance path Autopilot 𝑎𝑐 Control 𝛿𝑐 Aero_
𝑃𝑇 sensor processor system dynamics
𝑃𝑀
Missile
Position
Fundamentals of • Introduction to Missile Guidance
• Diff-Types of Homing Guidance
Navigation, Guidance • Active homing guidance
• Passive homing guidance
and Control • Types of Guidance Laws

A guided missile is defined as a space-


traversing unmanned vehicle which carries within itself the means for controlling
its flight path. A missile guidance system is defined as a group of components and
algorithms which measures the position of a guided missile with respect to its
target and causes changes in the flight path as required. The missile guidance
system includes sensing, computing, directing, stabilizing and servo-control
components. Guided missiles are primarily classified into different types (i.e.
categories) based on firing source and target such as:
• Surface-to-air • Air-to-surface • Surface-to-surface • Air-to-air

Guideline

Navigation in the context of missile guidance refers to the process of determining


and planning the path of the missile from its current position to a desired target
position. The missile needs to follow the planned path, which involves
continuously adjusting its heading and speed. This is where guidance algorithms
like vector field control come into play.

Guidance is a generic term that describes the hardware, the functions, and the
processes used to steer a missile to intercept a target. Therefore, we say that: a
Guidance is the process for guiding the path of an object towards a given point,
which in general may be moving. If the given point is fixed and the guided object is
manned (eg. migrating bird) then the process of determining and planning the path
is simply navigation. Thus, navigation can be said to be a subclass of guidance. A
guidance law is the algorithm by which the desired geometrical rule is
implemented. When the seeker (intercept error sensor) senses and determines the
instantaneous intercept error, the guidance processor then determines the
appropriate maneuver command, based on the guidance law, to reduce the error.
The autopilot in turn determines the control that is needed to achieve this
command and transmits the control signals to the control system actuators to
deflect the control surfaces.
Once the guidance processor has determined the magnitude and direction of the
error in the missile flight path and the autopilot has determined the steering
command, the missile control system must adjust the control surfaces to produce
the acceleration required to correct the flight path. This corrective acceleration is
applied in a lateral direction (perpendicular to the missile flight path) to change the
direction of the missile velocity vector. Missiles use tail, wing, canard, or thrusters
to maneuver and control their attitude. Some missiles use a combination of them.

Now, the question arises as to what input the actuator should apply to the missile.
This is the purpose of the control system, which is typically consists of three parts:

• Navigation - where am I? (Determining and planning the path)


• Guidance - where do I want to be? (Find a desired state to intercept target)
• Control - how do I get there? (Make the missile perform a desired task)

In a bit more detail, the navigation part determines the current position from
sensor measurements. The guidance part determines what the desired position is ?
and what the error in orientation is ? (i.e. based on the navigation part). The
control part determines what input should be applied to the missile to correct the
error, and sends this as a command to the actuators. A control system is the brain
of any system (aerospace, mechanical, electrical, mechatronic, chemical, etc.). The
purpose of a control system is to make the system perform a desired task. In order
to maintain such purpose, movable control surfaces are deflected by commands
from the guidance system in order to direct the missile, that is, the guidance
system will place the missile on the proper trajectory to intercept the target.
Target flight Missile
Intercept Heading
Position
error 𝑒 Guidance path Autopilot 𝑎𝑐 Control 𝛿𝑐 Aero_
𝑃𝑇 sensor processor system dynamics
𝑃𝑀
Missile
Position

A guided missile engagement is a highly dynamic process. The conditions that


determine how close the missile comes to the target are continuously changing,
sometimes at a very high rate. A guidance sensor measures one or more
parameters of the path of the missile relative to the target. A logical process is
needed to determine the required flight path corrections based on the sensor
measurements. This logical process is called a guidance law. The objective of a
guidance law is to cause the missile to come as close as possible to the target.
Guidance laws usually can be expressed in mathematical terms and are
implemented through a combination of electrical circuits and mechanical control
functions. A number of different schemes and their many variations have been
used for missile guidance, here we mention a few of them: pure pursuit, deviated
pursuit, parallel navigation, proportional navigation, beam-rider, and other
methods based on modem control theory.
There are many types of target interception, but here we explain graphically a few
of them, namely: pure pursuit, deviated pursuit and the parallel navigation

𝛿(𝑡) = 0 𝛿(𝑡) = 𝛼 ≠ 0 𝜆(𝑡) = 𝛼

𝜆 𝜆 𝜆

𝛿
𝐓𝐚𝐫𝐠𝐞𝐭 𝐓𝐚𝐫𝐠𝐞𝐭
𝐓𝐚𝐫𝐠𝐞𝐭
(A) Pure pursuit. (B) Deviated pursuit. (C) Parallel navigation.

𝛿(𝑡): deviation angle between the line-of-sight and the interceptor’s velocity vector.
𝜆(𝑡): The angle of line-of-sight to the reference line.

The simplest procedure for a guided missile directed at a moving target is to


describe a pursuit path. The missile is then headed along the line-of-sight at every
instant. Consequently, the rate of turn of the missile is always equal to the rate of
turn of the line-of-sight. Pursuit paths are highly curved near the end of flight.
(This is the pure pursuit path; other types of pursuit path will be discussed later.)

Target

Path of pursuit
Missile

Parallel navigation course: defined by the rule 𝜆(𝑡) = 𝜆(0), and leads incidentally to
a form of motion camouflage, because the pursuer (interceptor) appears stationary
against a distant background [also called constant absolute target direction or
constant bearing angle]. Parallel navigation describes the fact that as the two
objects approach each other on a collision course, the line-of-sight does not rotate
relative to the external world (here represented as ref-line). This means that the
bearing angle 𝜆, formed between ref-line and the LOS does not change over time.
This section is-intended to provide an overall
view of a guided missile and its various components. The principal components of
a guided missile are

Missile seeker and radome error analysis has been carried out extensively.
Basically, the main function of the missile seeker (also known as homing eye)
subsystem is to: (1) Provide the measurements of target motion required to
mechanize the guidance law. (2) Track the target with the antenna or other energy-
receiving device. (3) Track the target continuously after acquisition. (4) Measure
the 𝐿𝑂𝑆 (line-of-sight) angular rate 𝑑𝜆/𝑑𝑡. (5) Stabilize the seeker against a missile
pitching rate 𝑑𝜃𝑚 /𝑑t (also, yawing rate) that may be much larger than the LOS rate
𝑑𝜆/𝑑𝑡 to be measured. (6) Measure the closing velocity v𝑐 (i.e. this is possible with
some radars but difficult with IR seekers). The typical classical seeker hardware
consists of two/three gimbals on which are mounted gyroscopes and an antenna.

The warhead may be designed to inflict any of several possible kinds of


damage on the enemy. The warhead is the reason that the missile exists, the other
components are intended merely to ensure that the warhead will reach its
destination. A warhead is the section of a device that contains the explosive agent
or toxic (biological, chemical, or nuclear) material that is delivered by a missile,
rocket, torpedo, or bomb.

The airframe is the physical structure that carries the warhead to the
target, and contains the propulsion, guidance, and control system.

This system contains (Jet, Engines, Motors and


Thrusters) and provides the energy required to move the missile from the
launcher to the target.

The control system has two functions. It keeps the


missile in stable flight, and it translates the commands of the guidance
system into motion of the control surfaces, or into some other means for
modifying the missile trajectory.

This system determines whether the missile is on


the ordered trajectory of the ordered velocity. If the missile is off course,
the guidance system sends error signals to the control system.

The autopilot is a link between the function that indicates a change of


heading is needed (guidance processor) and the mechanism that can change the
heading (control system). The "autopilot" receives guidance commands and
processes them to the controls such as deflections or rates of deflection of control
surfaces or jet controls.

The designs of the actuators of different missiles may vary considerably,


but all have a common purpose, i.e., to convert autopilot commands into fin
deflections (i.e. fines and canards). The input to the model is the control--surface
deflection command; the output is the control-surface defection achieved.
A guidance system is called homing if
missile 𝑀 detects the target 𝑇 and tracks it thanks to energy emitted by the latter.
If the source of the energy is 𝑇 proper, e.g., radio transmissions, acoustic noise,
heat, the homing is called passive. If 𝑇 reflects energy beamed at it by 𝑀, then this
is active homing. In semi-active systems, 𝑀 homes onto energy reflected by 𝑇, the
latter being "illuminated" by a source I external to both. As stated here, homing
guidance may be of the active, semi-active, or passive type.
The expression homing guidance is used to describe a missile system that can
sense the target by some means, and then guide itself to the target by sending
commands to its own control surfaces. Homing is useful in tactical missiles where
considerations such as autonomous operation usually require sensing of target
motion to be done from the interceptor missile (or pursuer) itself.

are able to guide themselves independently after launch to


the target. These missiles are of the so-called launch-and-leave class. Therefore, an
active guided missile carries the radiation source. The radiation from the
interceptor missile is radiated, strikes the target, and is reflected back to the
missile. Thus, the missile guides itself on this reflected radiation. With active
homing, the missile contains both a radar transmitter and a receiver.

utilizes radiation originated by the target, or by some


other source not part of the overall weapon system. Typically, this radiation is in
infrared or the visible region, but may also occur in the microwave region.

uses a combination of active and passive guidance. A


source of radiation is part of the system, but is not carried in the missile. More
specifically, the independent source of radiation, radiates energy to the target,
whereby the energy is reflected back to the missile. As a result, the missile senses
the reflected radiation and homes on it.

Target
Reflected Radar Waves
Signals From Ship
Radar Waves
Reflected
From Missile
Signals

Missile Missile
Active Homing Semi-Passive Homing

3 Types of
Target

Missile
Homing Systems
Passive Homing
In 2017, Russian weapons manufacturer
"Tactical Missiles Corporation" announced that it was developing missiles that
would use artificial intelligence to choose their own targets. In 2019, the United
States Army announced it was developing a similar technology.

It is well-known that, missiles are so


important in modern warfare, and it will be of great importance when the targets
are dynamically movable objects. To take down a random moving target requires
ordnance that is smart enough to manoeuver, follow, chase and hopefully get close
enough to cause damage. However, the verbs "follow" and "chase" hide quite a lot
of subtly in implementation. Over the years, more and more sophisticated systems
have been developed to implement guidance control rules. In roughly chronological
order, and complexity are:

• Pure pursuit (or just pursuit) • Line of sight (Beam rider)


• Parallel navigation (Constant bearing) • Zero effort miss-distance
• Proportional navigation • Optimal navigations (State Space form)
• Modern navigations (Artificial intelligent based navigation)

The pursuit curve (also the radiodrome) was first studied by Pierre
Bouguer in 1732 and subsequently by the mathematician Boole. In an article on
navigation, Bouguer defined a curve of pursuit to explore the way in which one
ship might maneuver while pursuing another. In geometry, a curve of pursuit is a
curve constructed by analogy to having a point or points representing pursuers
and pursues, the curve of pursuit is the curve traced by the pursuers.

One of the most obvious and primitive guidance laws is pursuit guidance, in which
the missile velocity vector is directed toward the position of the target at any
instant in time. A variation of pursuit guidance that introduces the concept of
leading the target is deviated pursuit guidance. In this form of guidance, the angle
between the missile velocity vector and the line of sight to the target is held
constant. Both the pursuit and deviated pursuit guidance laws require a very high
missile turning rate close to the time of intercept.

The pure pursuit (simply pursuit) essentially means that the vector pointing from
the vehicle (i.e, missile) to the target, when scaled by the target’s speed, is aligned
with the target’s velocity vector scaled by the distance between the vehicle and the
target. This alignment is the core concept of the pure pursuit algorithm, ensuring
the vehicle steers towards the target. The criterion for the pure pursuit course is
that the missile always heads directly toward the present target position:

𝑑𝐫⃗𝑚 𝑑𝐫⃗𝑚 𝐯⃗⃗𝑚 𝐫⃗


‖𝐫⃗‖ =‖ ‖ 𝐫⃗ or = or ⃗⃗𝑟
𝐯⃗⃗𝑚 = v𝑚 𝐞 where ⃗⃗𝑟 = 𝐫⃗/𝑟
𝐞
𝑑𝑡 𝑑𝑡 v𝑚 𝑟
𝐫⃗𝑚 : The current position of the vehicle (or missile). 𝐫⃗𝑡 : The position of the target.
𝐫⃗ = 𝐫⃗𝑚 − 𝐫⃗𝑡 : The relative vector from missile to the target, with 𝑟 = ‖𝐫⃗‖ = ‖𝐫⃗𝑚 − 𝐫⃗𝑡 ‖.
𝐯𝑚 = 𝑑𝐫⃗𝑚 /𝑑𝑡: The velocity of the missile, with v𝑚 = ‖𝐯𝑚 ‖ = ‖𝑑𝐫⃗𝑚 /𝑑𝑡‖.
The equation ensures that the direction of the vehicle's motion (missile) aligns with
the direction of the target's future position. The missile should adjust its steering
so that its trajectory is directed towards where the target will be in the future,
rather than just where it currently is. To explain very well the pure pursuit
guidance we look for the following example: If a target 𝑇 moves along a known
curve, then a pursuers 𝑀 describes a pursuit curve if 𝑀 is always directed toward
𝑇 where: both 𝑇 and 𝑀 move with uniform velocities. Assume that the coordinates
of the point 𝑀 is 𝑴(𝑡) = (𝑥(𝑡), 𝑦(𝑡)) and its velocity is 𝐯𝑚 , then the speed of the
pursuers is given by v𝑚 = ‖𝑑𝐫𝑚 /𝑑𝑡‖. The vector connecting the target by the
pursuers is defined by 𝐫(𝑡) = (𝑥𝑡 − 𝑥, 𝑦𝑡 − 𝑦). The condition of tracking the target is
given by: "the velocity unit vector of 𝑀 should be equal to the unit connecting
vector", that is 𝐯⃗⃗𝑚 /v𝑚 = 𝐫⃗/𝑟.

The equation of target motion is given by 𝑻(𝑡) = (𝑥𝑡 , 𝑦𝑡 ) or 𝑻(𝑡) = 𝑥𝑡0 𝒆𝑥 + (𝑦𝑡0 + v𝑡 𝑡)𝒆𝑦
with 𝑥𝑡 = 𝑥𝑡0 and 𝑦𝑡 = 𝑦𝑡0 + v𝑡 𝑡. The point 𝑀 runs with the constant speed v𝑚
towards the instantaneous position of the target. The diff-equation corresponding
to 𝑴(𝑡) = (𝑥(𝑡), 𝑦(𝑡)), is consequently
1 𝑑 𝑥 𝑥𝑡 − 𝑥
𝑟. 𝐯⃗⃗𝑚 = v𝑚 . 𝐫⃗ ⟺ [(𝑥𝑡 − 𝑥)2 + (𝑦𝑡 − 𝑦)2 ]2 [𝑦] = v𝑚 [𝑦 − 𝑦]
𝑑𝑡 𝑡
𝑑𝑥 𝑑𝑦
= v𝑚 (𝑥𝑡 − 𝑥)[(𝑥𝑡 − 𝑥)2 + (𝑦𝑡 − 𝑦)2 ]−1/2 ; = v𝑚 (𝑦𝑡 − 𝑦)[(𝑥𝑡 − 𝑥)2 + (𝑦𝑡 − 𝑦)2 ]−1/2
𝑑𝑡 𝑑𝑡
It is possible to obtain closed-form analytic expression 𝑦 = f(𝑥) for the motion of 𝑀.
Notice that, if we let (𝑥𝑡 − 𝑥) ≥ 0 then:
1/2
𝑑𝑦 𝑦𝑡 − 𝑦 𝑑𝑥 v𝑚 𝑑𝑦𝑡 𝑑𝑦𝑡 𝑑𝑡 v𝑡 𝑑𝑦 2
= ; = ; with = = [1 + ( ) ]
𝑑𝑥 𝑥𝑡 − 𝑥 𝑑𝑡 [1 + (𝑑𝑦/𝑑𝑥)2 ]1/2 𝑑𝑥 𝑑𝑡 𝑑𝑥 v𝑚 𝑑𝑥
Multiplying both sides of the equation 𝑑𝑦/𝑑𝑥 = (𝑦𝑡 − 𝑦)/(𝑥𝑡 − 𝑥) by (𝑥𝑡 − 𝑥) and
taking the derivative with respect to 𝑥, we get
𝑑𝑦 𝑦𝑡 − 𝑦 𝑑𝑦 𝑑 𝑑 2 𝑦 𝑑𝑦 𝑑
= ⟺ 𝑦𝑡 − 𝑦 = (𝑥𝑡 − 𝑥) ⟺ (𝑦𝑡 − 𝑦) = (𝑥𝑡 − 𝑥) 2 + (𝑥 − 𝑥)
𝑑𝑥 𝑥𝑡 − 𝑥 𝑑𝑥 𝑑𝑥 𝑑𝑥 𝑑𝑥 𝑑𝑥 𝑡
1/2
v𝑡 𝑑𝑦 2 𝑑𝑦 𝑑 2 𝑦 𝑑𝑦
⟺ [1 + ( ) ] − = (𝑥𝑡 − 𝑥) 2 −
v𝑚 𝑑𝑥 𝑑𝑥 𝑑𝑥 𝑑𝑥
Therefore, we obtain following nonlinear second order diff-equation
2 1/2
v𝑚 2 𝑑2𝑦 𝑑𝑦 2 𝑑2 𝑦 v𝑡 𝑑𝑦 2
{( ) [𝑥𝑡 − 𝑥]} ( 2 ) = 1 + ( ) or [𝑥𝑡 − 𝑥] 2 = [1 + ( ) ]
v𝑡 𝑑𝑥 𝑑𝑥 𝑑𝑥 v𝑚 𝑑𝑥
From this relation, and let 𝑘 = v𝑡 /v𝑚 it follows that
𝑑 2 𝑦 𝑘√1 + (𝑑𝑦/𝑑𝑥)2 𝑑𝑦 1 𝐵 −𝑘
1 −𝐵
= ⇒ = sinh[𝐵 − 𝑘 ln(𝑥𝑡 − 𝑥)] = 𝑒 (𝑥𝑡 − 𝑥) − 𝑒 (𝑥𝑡 − 𝑥)𝑘
𝑑𝑥 2 (𝑥𝑡 − 𝑥) 𝑑𝑥 2 2
where 𝐵 is the constant of integration determined by the IC of 𝑦′ at time zero.
1 𝑒 −𝐵 𝑒𝐵
• If 𝑘 ≠ 1 or v𝑡 ≠ v𝑚 then 𝑦= { (𝑥𝑡 − 𝑥)1+𝑘 − (𝑥 − 𝑥)1−𝑘 } + 𝐶
2 1+𝑘 1−𝑘 𝑡
1 𝑒 −𝐵
• If 𝑘 = 1 or v𝑡 = v𝑚 then 𝑦= { (𝑥𝑡 − 𝑥)2 − 𝑒 𝐵 ln(𝑥𝑡 − 𝑥)} + 𝐶
2 2
The above system of diff-equation can be solved numerically using Euler method:
clear all, clc, Vm=3.5; Vt1=3; Vt2=1; dt=0.005; t=0; x0=1; y0=-10; M=[1.5;-8];
function dRm=pursuers(Vm,xt,yt,x,y)
D=sqrt((xt-x)^2+(yt-y)^2); dx=Vm*(xt-x)/D; dy=Vm*(yt-y)/D; dRm=[dx; dy];
end, e=1; T1= []; T2= []; X= []; Y= []; figure, hold on,
while e>0.01
xt = x0 + Vt1*t; yt = y0 + Vt2*t; x = M(1); y = M(2); % T and M positions
dRm = pursuers(Vm,xt,yt,x,y); e = sqrt((xt-x)^2+(yt-y)^2);
M = M + dt*dRm; % Euler Method
T1= [T1,xt]; T2= [T2,yt]; X= [X,x]; Y= [Y,y]; t = t+dt;
end, plot(X,Y,'b','LineWidth',1.5), grid on, plot(T1,T2,'r','LineWidth',1.5)

For a target that moves along a circle with center 𝑂 and radius 𝑅, we get the
differential system:
𝑑𝑥 v𝑚 𝑑𝑦 v𝑚
= (𝑅 cos 𝑡 − 𝑥); = (𝑅 sin 𝑡 − 𝑦); 𝐷 = [(𝑅 cos 𝑡 − 𝑥)2 + (𝑅 sin 𝑡 − 𝑦)2 ]1/2
𝑑𝑡 𝐷 𝑑𝑡 𝐷
The numerical solution of this system of diff-equation by using Euler method is:
clear all, clc, Vm=3.5; R=3; dt=0.005; t=0; M=[1.5; -8]; e=1;
function dRm=pursuers(Vm,xt,yt,x,y)
D=sqrt((xt-x)^2+(yt-y)^2); dx=Vm*(xt-x)/D; dy=Vm*(yt-y)/D; dRm=[dx; dy];
end, T1= []; T2= []; X= []; Y= []; figure, hold on,
while e>0.01
xt = R*cos(t); yt = R*sin(t); x = M(1); y = M(2);
dRm=pursuers(Vm,xt,yt,x,y); e = norm([xt-x,yt-y]);
M = M + dt*dRm; T1 = [T1,xt]; T2 = [T2,yt]; X = [X,x]; Y = [Y,y]; t = t+dt;
end, plot(X,Y,'b','LineWidth',1.5), grid on, plot(T1,T2,'r','LineWidth',1.5)

Form the analytical point of view, if a target, moves along the trajectory
(𝑅 cos(𝑡) , 𝑅 sin(𝑡)) then equations of motion for a pursuers passing by (0,0) is:
𝑘𝑅 𝑘𝑅
𝑥(𝑡) = 2
[𝑘 cos(𝑡) + sin(𝑡) − 𝑘𝑒 −𝑘𝑡 ]; 𝑦(𝑡) = [𝑘 sin(𝑡) − cos(𝑡) + 𝑘𝑒 −𝑘𝑡 ]
1+𝑘 1 + 𝑘2
The trajectory of the pursuers has the circle with center 𝑂 and radius 𝑘𝑅/(1 + 𝑘 2 )
as its asymptote, its speed goes to 𝑘𝑉/(1 + 𝑘 2 ), and its distance to the target goes
to 𝑘𝑅/(1 + 𝑘 2 ).
clear all, clc, dt=0.01; t=0:dt:10; R=3; k=7; Iter = length(t); figure;
for i=1:Iter
x(i)=(k*R/(1+k^2))*(k*cos(t(i))+sin(t(i))-k*exp(-k*t(i))); xt(i)=R*cos(t(i));
y(i)=(k*R/(1+k^2))*(k*sin(t(i))-cos(t(i))+k*exp(-k*t(i))); yt(i)=R*sin(t(i));
end
plot(xt,yt,'r','linewidth',1.5), grid on, hold on, plot(x,y,'b','linewidth',1.5)
It can be proved that the differential equation of the trajectory in spherical
coordinates (𝜃 = longitude, 𝜙 = colatitude, when target travels on the equator) is:
1/2
𝑑2𝜃 1 𝑑𝜃 2 𝑑𝜃 2
𝑑𝜃 2 𝑑𝜃 𝑐𝜙 𝑐𝜙 = cos 𝜙
= 𝑘 ( + ( ) 𝑐 𝑠
𝜙 𝜙 ) [1 + ( 𝑠 ) ] − ( ) 𝑐𝜙 𝑠𝜙 − 2 { };
𝑑𝜙 2 𝑐𝜙 𝑠𝜙 𝑑𝜙 𝑑𝜙 𝜙 𝑑𝜙 𝑑𝜙 𝑠𝜙 𝑠𝜙 = sin 𝜙

• Again a pure pursuit course is a course in which the missile velocity vector is
always directed toward the instantaneous target position. In deriving the equations
of motion (i.e. in polar form) the following assumptions are made:

(a) The target follows a non-maneuvering straight-line course.


(b) Both the missile and target speeds are constant.
(c) A two-dimensional coordinate system is set up

The geometry required for the derivation of the pure pursuit course equations of
motion in polar coordinates is given in figure, where the 𝐯⃗⃗𝑡 𝑇
following symbols are employed:
𝜃𝑡
𝑟: instantaneous distance 𝑀𝑇 between missile and target 𝐯⃗⃗𝑚 𝐫⃗
𝜃𝑡 : angle between target velocity vector and line of sight. 𝑀 𝜆=𝜃 𝑡

For an outgoing target the equations of motion are obtained by taking components
along 𝑟 and normal to 𝑟, 𝑟̇ = v𝑡 cos(𝜃𝑡 ) − v𝑚 ; and 𝑟𝜃̇𝑡 = −v𝑡 sin(𝜃𝑡 ). The solution of
these equations of motion is straightforward: 𝑟̇ /𝑟 = [𝑝 csc(𝜃𝑡 ) − cot(𝜃𝑡 )]𝜃̇𝑡 where
𝑝 = v𝑚 /v𝑡 . This can be integrated directly, giving
(sin(𝜃𝑡 ))𝑝−1 (1 + cos(𝜃0 ))𝑝 v𝑚
𝑟=𝐴 where for the initial values 𝑟0 and 𝜃0 : 𝐴 = 𝑟0 ; 𝑝=
(1 + cos(𝜃𝑡 ))𝑝 (sin(𝜃0 )) 𝑝−1 v𝑡
clear all, clc, dt=0.01; t=0; vm=17; vt=15; th0=pi/3; th=th0; xt= 8; yt= 3;
p=vm/vt; r0=16; A=r0*((1+cos(th0))^p)/(sin(th0))^(p-1); figure
%------------------------------------------------------------------------------%
for i=1:500
xt(i+1)= xt(i) - vt*dt; yt(i+1)= yt(i); % Target position
r(i)=A*((sin(th))^(p-1))/((1+cos(th))^p); % Relative position
x(i)= xt(i)-r(i)*cos(th); y(i)= yt(i)-r(i)*sin(th); % Missile position
th = th - (vt/r(i))*sin(th)*dt; % The angle of LOS
end
plot(xt,yt,'r','linewidth',1.5), grid on, hold on, plot(x,y,'b','linewidth',1.5)
For an incoming target, the target and missile motions both tend to decrease 𝑟 as
long as 𝜃𝑡 is acute. Here 𝜃𝑡 is monotonic increasing and hence the equations of
motion are 𝑟̇ = −v𝑡 cos(𝜃𝑡 ) − v𝑚 and 𝑟𝜃̇𝑡 = v𝑡 sin(𝜃𝑡 ). So, 𝑟̇ /𝑟 = −[𝑝 csc(𝜃𝑡 ) + cot(𝜃𝑡 )]𝜃̇𝑡
where 𝑝 = v𝑚 /v𝑡 . The solution by direct integration is
(1 + cos(𝜃𝑡 ))𝑝 (sin(𝜃0 ))𝑝+1
𝑟=𝐴 where for the initial values 𝑟0 and 𝜃0 : 𝐴 = 𝑟0
(sin(𝜃𝑡 ))𝑝+1 (1 + cos(𝜃0 ))𝑝
Note: A curve with multiple pursuers known as the mice problem, often creates a
beautiful pictures. The pursuers all start on a corner of a regular polygon, and
each pursues the point in the clockwise or counterclockwise direction (direction
must be uniform for all points). Usually points pursue the point directly next to
them, they can pursue points that are further away. The points will all eventually
meet in the center of the polygon.
The problem of pursuing a moving target is always one of the main topics in
navigation. In this part, we introduce a novel family of three dimensional pursuing
algorithms called accelerated pure pursuit problem.
Initiation: Given 𝐫⃗𝑚 (1), 𝐫⃗𝑡 (1), 𝐯⃗⃗𝑡 (𝑘), v𝑚 (1), and a𝑚
𝐫⃗(𝑘) = 𝐫⃗𝑚 (𝑘) − 𝐫⃗𝑡 (𝑘)
|| 𝐯⃗⃗𝑡 (𝑘) = [v𝑡𝑥 (𝑘); v𝑡𝑦 (𝑘); v𝑡𝑧 (𝑘)]
𝐫⃗𝑡 (𝑘 + 1) = 𝐫⃗𝑡 (𝑘) + 𝐯⃗⃗𝑡 . 𝑑𝑡
𝐯⃗⃗𝑚 (𝑘) = v𝑚 (𝑘)(𝐫⃗/‖𝐫⃗‖)
|| 𝐫⃗𝑚 (𝑘 + 1) = 𝐫⃗𝑚 (𝑘) + 𝐯⃗⃗𝑚 . 𝑑𝑡
v𝑚 (𝑘 + 1) = v𝑚 (𝑘) + a𝑚 . 𝑑𝑡

clear all, clc, dt=0.01; t=0; w=2; e=1; k=1; a=10; b=30; A0 = 5; V0 = 25;
Rt(:,1)=[0; 100; 100]; Rm(:,1)=[0; 100; 0]; figure;
while e>0.005
Rtm = Rt(:,k)-Rm(:,k); e = norm(Rtm);
Vt = [-a*cos(w*t); a*sin(w*t); b]; Rt(:,k+1) = Rt(:,k)+ Vt*dt;
Vm = V0*(Rtm./norm(Rtm)); Rm(:,k+1) = Rm(:,k)+ Vm*dt;
V0 = V0 + A0*dt; t = t + dt; k = k+1;
end
plot3(Rm(1,:), Rm(2,:), Rm(3,:),'b','linewidth',1.5), grid on, hold on,
plot3(Rt(1,:), Rt(2,:), Rt(3,:),'r','linewidth',1.5), grid on
the interceptor directs its velocity at a nonzero lead angle 𝛼
ahead of the line-of-sight, producing a deviated pursuit course defined by 𝛿(𝑡) = 𝛼.
𝐯𝑚 (𝑡) = v𝑚 (𝑡)[𝐫(𝑡)/𝑟(𝑡)] + v𝑚 (𝑡)𝐝; this is a deviated pursuit; with 𝐝 = constant vector
clear all, clc, dt=0.01; t=0; w=2; e=1; k=1; a=10; b=30; A0 = 5; V0 = 25;
Rt(:,1)=[0; 100; 100]; Rm(:,1)=[0; 100; 0]; d=[0.15; 0.17; 0.13]; figure;
while e>0.005
Rtm = Rt(:,k)-Rm(:,k); e = norm(Rtm);
Vt = [-a*cos(w*t); a*sin(w*t); b]; Rt(:,k+1) = Rt(:,k)+ Vt*dt;
Vm = V0*(Rtm./norm(Rtm)) + V0*d; Rm(:,k+1) = Rm(:,k)+ Vm*dt;
V0 = V0 + A0*dt; t = t + dt; k = k+1;
end
plot3(Rm(1,:), Rm(2,:), Rm(3,:),'b','linewidth',1.5), grid on, hold on,
plot3(Rt(1,:), Rt(2,:), Rt(3,:),'r','linewidth',1.5), grid on
𝛿(𝑡) = 𝛼 ≠ 0
𝐯𝑚 /v𝑚
𝐝
𝛿 𝐯⃗⃗𝑡
Deviated pursuit
𝐮LOS = 𝐫(𝑡)/𝑟(𝑡)
𝜃𝑡
𝛿
𝑇

𝛿 𝐯⃗⃗𝑚 𝐫⃗
𝛿
𝐓𝐚𝐫𝐠𝐞𝐭 𝛿(𝑡) = 𝛼
𝑀 𝜆
From the above vector diagram we have to write: 𝑟̇ (𝑡) = v𝑡 (𝑡) cos(𝜃𝑡 ) − v𝑚 (𝑡) cos(𝛿)
and 𝜆̇(𝑡) = [v𝑡 (𝑡) sin(𝜃𝑡 ) − v𝑚 (𝑡) sin(𝛿)]/𝑟(𝑡). These two statements are equivalent to
𝐯𝑚 (𝑡) = v𝑚 (𝑡)[𝐫(𝑡)/𝑟(𝑡)] + v𝑚 (𝑡)𝐝, but only in the 2-dimensional space. In case of
horizontally moving target we have 𝜆̇(𝑡) = −𝜃̇𝑡 (𝑡). The solution of these equations of
motion is straightforward:
𝑟(𝑘 + 1) = 𝑟(𝑘) + [v𝑡 cos(𝜃𝑡 ) − v𝑚 cos(𝛿)]𝑑𝑡
𝜃𝑡 (𝑘 + 1) = 𝜃𝑡 (𝑘) − [(v𝑡 /𝑟(𝑡)) sin(𝜃𝑡 ) − (v𝑚 /𝑟(𝑡)) sin(𝛿)]𝑑𝑡
clear all, clc, dt=0.01; vm=17; vt=15; th=pi/3; r=16; xt=8; yt=3; del=pi/9;
i=1; e=1; figure
while e>0.01
xt(i+1)= xt(i) - vt*dt; yt(i+1)= yt(i); % Target position
r(i+1)= r(i) + (vt*cos(th)-vm*cos(del))*dt; % Relative position
x(i)= xt(i)-r(i)*cos(th); y(i)= yt(i)-r(i)*sin(th); % Missile position
th = th - ((vt/r(i))*sin(th)-(vm/r(i))*sin(del))*dt; % The angle of LOS
e = norm(r(i)); i=i+1;
end
plot(xt,yt,'r','linewidth',1.5), grid on, hold on, plot(x,y,'b','linewidth',1.5)

The algorithm is based on a very simple idea, at every


instant the algorithm selects a point on the path called the "look-ahead point,"
which is a certain distance ahead of the current position of the interceptor. Then it
sets the value of the steering angle 𝛿(𝑡) to be equal and opposite to the offset error,
until the interceptor is directed towards its moving goal and its trajectory
converges with the path. The missile’s current position is represented by
coordinates (𝑥, 𝑦), and its orientation (heading) is given by an angle 𝛼𝑚 .
𝑅 = The missile’s turning radius (radius of curvature)
𝛿 = The steering angle and 𝐯𝑚 is the heading direction 𝑅 Path
𝛼 = The angle between 𝐯𝑚 and line of sight 𝛿
𝛼𝑚 = Angle between heading direction and center line 𝑅 (𝑥𝑡 , 𝑦𝑡 )
2𝛼 ℓ𝑑
ℓ𝑑 = look-ahead distance (m)
𝐿 = Missile length (m)
∆𝑦
𝑥𝑡 , 𝑥 = The 𝑥-components of both target and missile (m)
𝛼 𝐯𝑚
𝑦𝑡 , 𝑦 = The 𝑦-components of both target and missile (m) 𝐿
∆𝑥 = 𝑥𝑡 − 𝑥 (m) The relative 𝑥-distance (𝑥, 𝑦) 𝛼𝑚
∆𝑥
∆𝑦 = 𝑦𝑡 − 𝑦 (m) The relative 𝑦-distance
The missile’s steering angle 𝛿 can be defined using the goal point location and the
angle 𝛼 between the missile’s heading vector and the look-ahead vector. By
applying the law of sines, it is possible to derive the following equation
ℓ𝑑 𝑅 ℓ𝑑 𝑅 ℓ𝑑
= ⟺ = ⟺ =𝑅
sin(2𝛼) sin(𝜋/2 − 𝛼) 2 sin(𝛼) cos(𝛼) cos(𝛼) 2 sin(𝛼)
The curvature of the circular arc is given by 𝜅 = 1/𝑅 therefore, 𝜅 = 2 sin(𝛼) /ℓ𝑑 . The
steering angle can be written as follows: tan(𝛿) = 𝐿/𝑅 ⟹ 𝛿 = tan−1(𝜅𝐿). The Look-
Ahead Pursuit control law can be expressed as: 𝛿 = tan−1(2𝐿 sin(𝛼) /ℓ𝑑 ). If we define
𝛽 = 𝛼 + 𝛼𝑚 then tan(𝛽) = ∆𝑦/∆𝑥 means that 𝛼 = tan−1(∆𝑦/∆𝑥) − 𝛼𝑚 . Now, we use the
fact that v𝑚 = 𝑅(𝑑𝛼𝑚 /𝑑𝑡) so we have 𝑑𝛼𝑚 /𝑑𝑡 = (v𝑚 /𝑅) = (v𝑚 /𝐿)(𝐿/𝑅) = (v𝑚 /𝐿) tan(𝛿)

clc; clear; close all; clf;


% ------------------------------------------------------------------------------
dt = 0.1; % Time step (s)
L = 2.5; % Missile length (m)
LD = 5; % Look-ahead distance (m)
N = 70; % Number of simulation steps
vm = 16.5; % Missile speed (m/s)
vt = 17; % Target speed (m/s)
x = 0; y = -20; alpham = 0; % Initial states of the missile
Pt = [17, 0]; Pt = zeros(1, 2); t=0;
% ------------------------------------------------------------------------------
for i = 1:N
Dir = [cos(-pi*t/4), sin(-pi*t/4)]; % Target direction
Pt(i+1,:) = Pt(i,:) + vt*Dir*dt; % dPt/dt= [vt*cos(th),vt*sin(th)]
xt(i) = Pt(i,1); yt(i) = Pt(i,2); del_y=yt(i)-y(i); del_x=xt(i)-x(i);
alpha = atan2(del_y, del_x) - alpham; % tan(beta)=del(y)/del(x)
delta = atan2(2*L*sin(alpha), LD); % tan(delta)=L/R
x(i+1) = x(i) + vm*cos(alpham)*dt; % vm*cos(alpham)=dx/dt
y(i+1) = y(i) + vm*sin(alpham)*dt; % vm*sin(alpham)=dy/dt
alpham = alpham + (vm/L)*tan(delta)*dt; % (vm/L)*tan(delta)=dalpham/dt
Pm(i, :) = [x(i), y(i)]; t=t+dt;
end
plot(Pt(:, 1), Pt(:, 2), 'r--', 'LineWidth', 2); hold on;
plot(Pm(:, 1), Pm(:, 2), 'b-', 'LineWidth', 2);
plot(x(end), y(end), 'go', 'MarkerSize', 10, 'MarkerFaceColor', 'g');
plot(xt(end), yt(end), 'mo', 'MarkerSize', 10, 'MarkerFaceColor', 'm');
xlabel('X'); ylabel('Y'); grid on; title('Missile Tracking a Moving Target');
The constant bearing course is followed
by a missile when the line-of-sight "LOS" from missile to target maintain constant
direction in space, i.e. 𝜆 constant. This is often referred to as a collision course.
Since a collision is the ultimate result in all the flight paths discussed, the term
constant bearing course is more descriptive. It is evident that the missile velocity
must be greater than the target velocity.

Let we define 𝐯𝑚 = missile speed, 𝐯𝑡 = target speed, and ‖𝐫⃗‖ = relative distance. The
parallel navigation rule 𝜆̇(𝑡) = 0 (or 𝜔 = 0) is given by where 𝜆 is the angle of line-of-
sight "LOS" with some reference line. According to this rule the direction of LOS is
kept constant relative to inertial frame [i.e. the LOS is kept parallel to initial LOS]

In three-dimensional vector geometry, the parallel navigation rule is given by 𝝎 ⃗⃗


⃗⃗⃗⃗ = 𝟎
where 𝝎⃗⃗⃗⃗ is the rate of rotation of the line-of-sight 𝐫⃗ = 𝐫⃗𝑡 − 𝐫⃗𝑚 . Let we define the
following vectors 𝐯⃗⃗𝑡𝑚 = 𝑑𝐫⃗/𝑑𝑡 = 𝐯⃗⃗𝑡 − 𝐯⃗⃗𝑚 = 𝝎
⃗⃗⃗⃗ × 𝐫⃗ and 𝐯⃗⃗𝑐 = −𝐯⃗⃗𝑡𝑚 = 𝐫⃗ × 𝝎
⃗⃗⃗⃗ then
⃗⃗
⃗⃗⃗⃗ = 𝟎
𝝎 ⟺ ⃗⃗, because 𝐫⃗ × 𝐯⃗⃗𝑐 = 𝐫⃗ × (𝐫⃗ × 𝝎
𝐫⃗ × 𝐯⃗⃗𝑐 = 𝟎 ⃗⃗⃗⃗) = 𝑟 2 𝝎 ⃗⃗
⃗⃗⃗⃗ = 𝟎
The vector geometry give
𝑑𝑟
= v𝑡 (𝑡) cos(𝜃𝑡 ) − v𝑚 (𝑡) cos(𝜃𝑚 ) 𝐯⃗⃗𝑡
𝑑𝑡
𝑑𝜆 𝜃𝑡
𝑟 = [v𝑡 (𝑡) sin(𝜃𝑡 ) − v𝑚 (𝑡) sin(𝜃𝑚 )] = 0 𝑇
𝑑𝑡
𝐯⃗⃗𝑚
The parallel navigation rule implies that
𝜆̇(𝑡) = 0 ⟹ 𝜔(𝑡) = 0 ⟹ v𝑡 (𝑡) sin(𝜃𝑡 ) = v𝑚 (𝑡) sin(𝜃𝑚 ) 𝜃𝑚 𝐫⃗
v𝑚 𝑀 𝜆
𝜃𝑡 = asin ( sin(𝜃𝑚 )) This is the collision triangle rule
v𝑡
The goal of the missile guidance system is to steer a missile heading to collide with
a target. The collision theory goes back to the old sailor’s saying, which predicts
the collision with other ships. This is called, ‘Constant Bearing, Decreasing Range’.
The collision triangle is defined as the triangle formed by the initial positions of
missile and target, and the intercept point where the missile hits the target when
flown in a straight line.
𝛽 = 𝜋 − 𝜃𝑡 The law of sine gives
v𝑚 v𝑚 v𝑚 v𝑡
𝐯⃗⃗𝑚 𝐯⃗⃗𝑡 = = =
sin(𝛽) sin(𝜋 − 𝜃𝑡 ) sin(𝜃𝑡 ) sin(𝜃𝑚 )
𝜃𝑡 Means that
𝛽
𝜃𝑚 v𝑡 v𝑚
𝜃𝑚 = sin−1 ( sin(𝜃𝑡 )) or 𝜃𝑡 = sin−1 ( sin(𝜃𝑚 ))
𝐯⃗⃗𝑡𝑚 v𝑚 v𝑡
The first condition of the collision triangle is the LOS angle 𝜆 is constant.
This is same as the rate of change of the LOS is zero. The second condition of the
collision triangle is the LOS distance 𝐫⃗ = 𝐫⃗𝑡 − 𝐫⃗𝑚 should decrease. These two
conditions can be expressed mathematically, 𝜆 = constant and 𝑟̇ < 0.
v𝑚
𝜃𝑡 = sin−1 ( sin(𝜃𝑚 )) and v𝑡 (𝑡) cos(𝜃𝑡 ) − v𝑚 (𝑡) cos(𝜃𝑚 ) < 0
v𝑡
LOS4
𝐯⃗⃗𝑚
𝐯⃗⃗𝑡 LOS3

𝜃𝑡 𝐯⃗⃗𝑚 LOS2
𝛿 LOS
𝜃𝑚 = 𝛿 𝜃𝑡
𝐯⃗⃗𝑐 𝜆 𝜆 𝜆 𝜆
LOS1
If we assume that v𝑡 (𝑡) = ‖𝐯⃗⃗𝑡 ‖ is variable and 𝜃𝑡 is constant then we have

𝑑v𝑡 𝑑𝜃𝑚 𝑑𝜃𝑚 𝑑v𝑡 sin(𝜃𝑡 )


v𝑡 (𝑡) sin(𝜃𝑡 ) = v𝑚 sin(𝜃𝑚 ) ⟺ sin(𝜃𝑡 ) = v𝑚 cos(𝜃𝑚 ) so v𝑚 =
𝑑𝑡 𝑑𝑡 𝑑𝑡 𝑑𝑡 cos(𝜃𝑚 )

Knowing that the normal acceleration of the missile is given by a𝑛 = v𝑚 (𝑑𝜃𝑚 /𝑑𝑡)
2
[because 𝐚⃗⃗𝑛 = (v𝑚 ⃗⃗ = v𝑚 (v𝑚 /𝑟)𝐧
/𝑟)𝐧 ⃗⃗ = (v𝑚 𝜔𝑚 )𝐧
⃗⃗ ⟹ a𝑛 = v𝑚 𝜔𝑚 = v𝑚 (𝑑𝜃𝑚 /𝑑𝑡)]. Notice
that v𝑡 sin(𝜃𝑡 ) = v𝑚 sin(𝜃𝑚 ) ⟺ v𝑚 [1 − cos (𝜃𝑚 )] = v𝑡2 sin2 (𝜃𝑡 ) so
2 2

𝑑𝜃𝑚 𝑑v𝑡 sin(𝜃𝑡 ) 𝑑v𝑡 𝑘 sin(𝜃𝑡 ) v𝑚


a𝑛 = v𝑚 = = ; 𝑘= ; cos(𝜃𝑚 ) = 𝑘/√𝑘 2 − sin2 (𝜃𝑡 )
𝑑𝑡 𝑑𝑡 cos(𝜃𝑚 ) 𝑑𝑡 √𝑘 2 − sin2 (𝜃𝑡 ) v𝑡

An alternative formula can be obtained if we use tan(𝜃𝑚 ) = sin(𝜃𝑡 ) /√𝑘 2 − sin2 (𝜃𝑡 )

𝑑v𝑡 sin(𝜃𝑡 ) 𝑑v𝑡


a𝑛 = 𝑘 =𝑘 tan(𝜃𝑚 )
𝑑𝑡 √𝑘 2 − sin2 (𝜃𝑡 ) 𝑑𝑡

The missile travels in such a way that its own rate of


turn is proportional to the rate of turn of the line-of-sight from missile to target at
any instant 𝛾̇ (𝑡) = 𝑁𝜆̇(𝑡). If the missile rate of
turn equals the line-of-sight rate of turn, the
flight path becomes a pursuit path (which is
really a special case of the proportional
navigation path). In general, in proportional
navigation, the missile rate of turn is a fixed
multiple of the line-of-sight rate paths are less
curved than pursuit paths, but more curved
than collision paths.

The guidance systems of most modern missiles


use a guidance law called proportional
navigation, in which turning is commanded at
a rate 𝛾̇ proportional to the rotation rate of the line-of-sight (i.e., the line-of-sight
rate, 𝜆̇), such that 𝛾̇ (𝑡) = 𝑁𝜆̇(𝑡) where 𝑁 is termed the navigation constant and falls
on the interval 3 ≤ 𝑁 ≤ 5 in missiles. (Note that PN refers specifically to the
guidance law stated in 𝛾̇ (𝑡) = 𝑁𝜆̇(𝑡) and is distinct from proportional control.)
Essentially, the PN rotates the missile heading at a rate that is proportional to the
LOS rate (means that 𝛾̇ (𝑡) = 𝑁𝜆̇(𝑡)). If the PN works perfectly, the LOS rate should
converge and stay at zero until interception. The kinematics tell us that the
commanded acceleration is a𝑐 = 𝑁v𝑐 𝜆̇(𝑡).
Explanation: In mechanics the normal acceleration is 𝛾 = 𝜃𝑚
given by 𝐚⃗⃗𝑛 = (v 2 /𝑟)𝐧
⃗⃗ = v𝜔𝐧
⃗⃗ ⟹ a𝑐 = v(𝑑𝜃/𝑑𝑡). So 𝑇 𝜃𝑡
𝑑𝜃𝑚 𝑑𝛾 𝑑𝜆 𝛾̇ = 𝑁𝜆̇ 𝐯⃗⃗𝑡
a𝑐 = v𝑐 = v𝑐 = 𝑁v𝑐 or simply a𝑐 = 𝑁v𝑐 𝜆̇(𝑡)
𝑑𝑡 𝑑𝑡 𝑑𝑡
In vector form we have
𝐫⃗ 𝜆̇
𝐫⃗ × 𝐯⃗⃗𝑐 𝐫⃗ × 𝐯⃗⃗𝑐
𝐚⃗⃗𝑐 = 𝑁𝝎
⃗⃗⃗⃗ × 𝐯⃗⃗𝑐 where ⃗⃗⃗⃗ =
𝝎 =
𝐫⃗. 𝐫⃗ ‖𝐫⃗‖2 𝐚⃗⃗𝑐 𝛿 𝐯⃗⃗𝑚
𝑁 𝛾
Finally we have 𝐚⃗⃗𝑐 = −𝑁𝐯⃗⃗𝑐 × ⃗𝝎
⃗⃗⃗ or 𝐚⃗⃗𝑐 = (𝐫⃗ × 𝐯⃗⃗𝑐 ) × 𝐯⃗⃗𝑐
‖𝐫⃗‖2 𝑀
Theoretically, the proportional navigation guidance law issues acceleration
commands, perpendicular to the instantaneous missile-target line-of-sight, which
are proportional to the line-of-sight rate and closing velocity. In tactical radar
homing missiles using proportional navigation guidance, the seeker provides an
effective measurement of the line-of-sight rate, and a Doppler radar provides
closing velocity information. In tactical IR missile applications of proportional
navigation guidance, the line-of-sight rate is measured, whereas the closing
velocity, required by the guidance law, is "guesstimated." To better understand
how proportional navigation works, let us consider the two dimensional, point
mass missile-target engagement geometry.

We can see from the figure that the missile, with velocity magnitude 𝐯𝑚 , is heading
at an angle of 𝜃 = 𝜃Lead + 𝜃He with
respect to the line of sight. The angle 𝜃 = 𝜃Lead + 𝜃He 𝐯𝑡 𝐚𝑡
𝜃Lead is known as the missile lead angle.
The lead angle is the theoretically correct
𝐯𝑚 𝛽
angle for the missile to be on a collision
triangle with the target. In other words, if Target
the missile is on a collision triangle, no 𝐚𝑐
𝐫
further acceleration commands are 𝜃
required for the missile to hit the target. 𝜆
The angle 𝜃He is known as the heading
Missile
error. This angle represents the initial
deviation of the missile from the collision triangle. In figure the imaginary line
connecting the missile and target is known as the line of sight. The line of sight
makes an angle of 𝜆 with respect to the fixed reference, and the length of the line of
sight (instantaneous separation between missile and target) is a range denoted 𝐫.
From a guidance point of view, we desire to make the range between missile and
target at the expected intercept time as small as possible (hopefully zero). The
point of closest approach of the missile and target is known as the miss distance.
The closing velocity v𝑐 is defined as the negative rate of change of the distance from
the missile to the target, or v𝑐 = −𝑑‖𝐫⃗‖/𝑑𝑡. Therefore, at the end of the
engagement, when the missile and target are in closest proximity, the sign of the
closing velocity will change. In other words, from calculus we know that the closing
velocity will be zero when 𝑟 = ‖𝐫⃗‖ is a minimum (that is, the function is either
minimum or maximum when its derivative is zero). The desired acceleration
command 𝐚⃗⃗𝑐 , which is derived from the proportional navigation guidance law, is
perpendicular to the instantaneous line of sight. In our engagement model, the
target can maneuver evasively with acceleration magnitude 𝐚⃗⃗𝑡 . Since target
acceleration 𝐚⃗⃗𝑡 in the preceding model is perpendicular to the target velocity vector,
the angular velocity of the target can be expressed as

𝛽(𝑡) = (a𝑡 /v𝑡 )𝑡 ⟺ 𝑑𝛽/𝑑𝑡 = (a𝑡 /v𝑡 ) ⟺ 𝛽(𝑡) = 𝛽(𝑡 − ∆𝑡) + (a𝑡 /v𝑡 )∆𝑡

where v𝑡 is the magnitude of the target velocity. The components of the target
velocity vector in the Earth or inertial coordinate system can be found by
integrating the differential equation given earlier for the flight-path angle of the
target 𝛽(𝑡) and substituting in v𝑡𝑥 (𝑡) = −v𝑡 cos 𝛽(𝑡) and v𝑡𝑦 (𝑡) = v𝑡 sin 𝛽(𝑡). Target
position components in the Earth fixed coordinate system can be found by directly
integrating the target velocity components. Therefore, the differential equations for
the components of the target position are given by
v𝑡𝑥 𝑃𝑡𝑥
𝐫̇𝑡 (𝑡) = 𝐯𝑡 ⟺ 𝐫𝑡 (𝑡) = 𝐫𝑡 (𝑡 − ∆𝑡) + 𝐯𝑡 ∆𝑡 with 𝐯𝑡 (𝑡) = [v ] ; and 𝐫𝑡 (𝑡) = [𝑃 ]
𝑡𝑦 𝑡𝑦

Similarly, the missile velocity and position differential equations are given by
𝐯̇ 𝑚 (𝑡) = 𝐚𝑚 ⟺ 𝐯𝑚 (𝑡) = 𝐯𝑚 (𝑡 − ∆𝑡) + 𝐚𝑚 ∆𝑡 v𝑚𝑥 𝑃𝑚𝑥
with 𝐯𝑚 (𝑡) = [v ] and 𝐫𝑚 (𝑡) = [𝑃 ]
𝐫̇𝑚 (𝑡) = 𝐯𝑚 (𝑡) ⟺ 𝐫𝑚 (𝑡) = 𝐫𝑚 (𝑡 − ∆𝑡) + 𝐯𝑚 ∆𝑡 𝑚𝑦 𝑚𝑦
𝑇
Where 𝐚𝑚 = [a𝑚𝑥 , a𝑚𝑦 ] and a𝑚𝑥 and a𝑚𝑦 are the missile acceleration components
in the Earth coordinate system. To find the missile acceleration components, we
must first find the components of the relative missile-target separation. This is
accomplished by first defining the relative missile-target separations
𝑟𝑥
𝐫(𝑡) = 𝐫𝑡 (𝑡) − 𝐫𝑚 (𝑡) with 𝐫(𝑡) = [𝑟 ]
𝑦

We can see from the above figure that the line-of-sight angle can be found, using
trigonometry, in terms of the relative separation components as 𝜆 = tan−1 (𝑟𝑦 /𝑟𝑥 ). If
we define the relative velocity components in Earth coordinates to be
v𝑡𝑚𝑥
𝐯𝑡𝑚 (𝑡) = 𝐯𝑡 (𝑡) − 𝐯𝑚 (𝑡) with 𝐯𝑡𝑚 (𝑡) = [v ]
𝑡𝑚𝑦

we can calculate the line-of-sight rate by direct differentiation of the expression for
line-of-sight angle. The expression for the line-of-sight rate 𝜆̇ is given by

𝑑𝜆 1 𝑑 𝑑𝑢/𝑑𝑡
= 2 (𝑟𝑥 v𝑡𝑚𝑦 − 𝑟𝑦 v𝑡𝑚𝑥 ) by using tan−1 (𝑢(𝑡)) = and 𝑟 2 = 𝑟𝑥2 + 𝑟𝑦2
𝑑𝑡 𝑟 𝑑𝑡 1 + 𝑢2
Because the closing velocity is defined as the negative rate of change of the missile
target separation, it can be obtained by differentiating the preceding equation,
yielding v𝑐 = −𝑑𝑟/𝑑𝑡 = −𝑑(√𝑟𝑥2 + 𝑟𝑦2 )/𝑑𝑡 = −(𝑟𝑥 v𝑡𝑚𝑥 + 𝑟𝑦 v𝑡𝑚𝑦 )/𝑟
𝑑𝜆 1 𝑑𝑟 1
Finally we have = 2 (𝑟𝑥 v𝑡𝑚𝑦 − 𝑟𝑦 v𝑡𝑚𝑥 ); and v𝑐 = − = − (𝑟𝑥 v𝑡𝑚𝑥 + 𝑟𝑦 v𝑡𝑚𝑦 )
𝑑𝑡 𝑟 𝑑𝑡 𝑟
The magnitude of the missile guidance command a𝑐 can then be found from the
definition of proportional navigation, or a𝑐 = 𝑁v𝑐 𝜆̇. Because the acceleration
command is perpendicular to the instantaneous line of sight, the missile
acceleration components in Earth coordinates can be found by trigonometry. The
missile acceleration components are
a𝑚𝑥 = −a𝑐 sin 𝜆 − sin 𝜆
or 𝐚𝑚 = a𝑐 𝐮𝑐 with 𝐮𝑐 = [ ]
a𝑚𝑦 = a𝑐 cos 𝜆 cos 𝜆
We have now listed all of the differential equations required to model a complete
missile-target engagement in two dimensions. However, some additional equations
are required for the initial conditions on the differential equations in order to
complete the engagement model.

A missile employing proportional navigation guidance is not fired at the target but
is fired in a direction to lead the target. The initial angle of the missile velocity
vector with respect to the line of sight is known as the missile lead angle 𝜃Lead . The
theoretical missile lead angle can be found by application of the law of sines,
yielding 𝜃Lead = sin−1((v𝑡 /v𝑚 ) sin(𝛽 + 𝜆)). In practice, the missile is usually not
launched exactly on a collision triangle, as the expected intercept point is not
known precisely. The location of the intercept point can only be approximated
because we do not know in advance what the target will do in the future. In fact,
that is why a guidance system is required! Any initial angular deviation of the
missile from the collision triangle is known as a heading error 𝜃He . The initial
missile velocity components can therefore be expressed in terms of the lead angle
𝜃Lead and actual heading error 𝜃He as v𝑚𝑥 (0) = v𝑚 cos(𝜃) and v𝑚𝑦 (0) = v𝑚 sin(𝜃)
where 𝜃 = 𝜃Lead + 𝜃He . To witness and understand the effectiveness of proportional
navigation, it is best to simulate the guidance law and test its properties under a
variety of circumstances.

𝐫 = 𝐫𝑡 − 𝐫𝑚 and 𝐯𝑡𝑚 = 𝐯𝑡 − 𝐯𝑚 𝐔𝐩𝐝𝐚𝐭𝐞𝐬 𝐛𝐲 𝐄𝐮𝐥𝐞𝐫 𝐦𝐞𝐭𝐡𝐨𝐝


v𝑐 = −(𝑟𝑥 v𝑡𝑚𝑥 + 𝑟𝑦 v𝑡𝑚𝑦 )/𝑟 a𝑛𝑡
| 𝛽(𝑘 + 1) = 𝛽(𝑘) + ( ) . 𝑑𝑡
𝜆 = tan−1 (𝑟𝑦 /𝑟𝑥 ) v𝑡
𝑑𝜆 1 𝐫𝑡 (𝑘 + 1) = 𝐫𝑡 (𝑘) + 𝐯𝑡 . 𝑑𝑡
= 2 (𝑟𝑥 v𝑡𝑚𝑦 − 𝑟𝑦 v𝑡𝑚𝑥 ) | 𝐫𝑚 (𝑘 + 1) = 𝐫𝑚 (𝑘) + 𝐯𝑚 . 𝑑𝑡
𝑑𝑡 𝑟
a𝑐 = 𝑁v𝑐 (𝑑𝜆/𝑑𝑡) 𝐯𝑚 (𝑘 + 1) = 𝐯𝑚 (𝑘) + 𝐚𝑚 . 𝑑𝑡
− sin 𝜆 |
𝐚𝑚 = a𝑐 𝐮𝑐 with 𝐮𝑐 = [ ] 𝐍𝐨𝐭𝐞:
cos 𝜆 If there is no convergence we use other update method
− cos 𝛽 | such as the or the 2end order Runge– Kutta or even the
𝐯𝑡 = v𝑡 𝐮𝑡 with 𝐮𝑡 = [ ]
sin 𝛽
𝑑𝛽/𝑑𝑡 = a𝑛𝑡 /v𝑡 4th order Runge– Kutta method.
%------------------------------------------------------------------
% Simulation of Tactical Missiles in 2D (Proportional Navigation):
%------------------------------------------------------------------
clear all, clc,
%------------------------------------------------------------------
Hedeg=-20; % Initial heading error angle in degree
Beta=-30; % Initial orientation of target
N=4; % Proportional navigation constant
vm=300; vt=100; % Initial velocities of missile and target
Xnt=9.66; % Normal acceleration (target)
Rm=[0; 1000]; Rt=[4000; 1000]; % Initial positions
Vt = vt.*[-cos(Beta); sin(Beta)]; % target velocity
He=Hedeg/57.3; Rtm=Rt-Rm; R=norm(Rtm); % relative position
Xlam=atan2(Rtm(2),Rtm(1)); % line of sight angle
Xlead=asin(vt*sin(Beta+Xlam)/vm); % the missile lead angle
Thet = Xlam + Xlead;
Vm = vm.*[cos(Thet+He); sin(Thet+He)]; % missile velocity
Vtm=Vt-Vm; % relative velocity
Vc=-(Rtm(1)*Vtm(1)+ Rtm(2)*Vtm(2))/R; % closing speed
%------------------------------------------------------------------
dt=0.001; i=1; t=0;
while Vc>=0.01,
Rtm = Rt(:,i)-Rm(:,i); Vtm = Vt-Vm; R = norm(Rtm);
Vc = -(Rtm(1)*Vtm(1)+ Rtm(2)*Vtm(2))/R; % v𝑐 closing speed
Xlam = atan2(Rtm(2),Rtm(1)); % 𝜆 LOS angle
Xlamd = (Rtm(1)*Vtm(2) - Rtm(2)*Vtm(1))/R; % 𝑑𝜆/𝑑𝑡 LOS rate
Xnc = N*Vc*Xlamd; % a𝑐 = 𝑁v𝑐 (𝑑𝜆/𝑑𝑡)
Am=Xnc.*[-sin(Xlam); cos(Xlam)]; % a𝑚 = a𝑐 𝐮𝑐
Vt = vt.*[-cos(Beta); sin(Beta)]; % 𝐯𝑡 = v𝑡 𝐮𝑡
dB=Xnt/vt; % 𝑑𝛽/𝑑𝑡 = a𝑛𝑡 /v𝑡
%---------------- Updates by Euler methods -----------------
Beta = Beta + dt*dB; % 𝛽(𝑘 + 1) = 𝛽(𝑘) + (a𝑛𝑡 /v𝑡 ). 𝑑𝑡
Rt(:,i+1) = Rt(:,i) + dt*Vt; % 𝐫𝑡 (𝑘 + 1) = 𝐫𝑡 (𝑘) + 𝐯𝑡 . 𝑑𝑡
Rm(:,i+1) = Rm(:,i) + dt*Vm; % 𝐫𝑚 (𝑘 + 1) = 𝐫𝑚 (𝑘) + 𝐯𝑚 . 𝑑𝑡
Vm = Vm + dt*Am; t=t+dt; i=i+1; % 𝐯𝑚 (𝑘 + 1) = 𝐯𝑚 (𝑘) + 𝐚𝑚 . 𝑑𝑡
end
%------------------------------------------------------------------
Range = R, figure, hold on, plot(Rt(1,:), Rt(2,:),'linewidth',1.5),
grid on, plot(Rm(1,:), Rm(2,:),'linewidth',1.5), hold off
title('Two-dimensional tactical missile-target engagement')
xlabel('Downrange (m)'), ylabel('Altitude (m)')
%------------------------------------------------------------------

In tactical missile-target engagement the diff-equations of the missile and target


could be solved using 2end or 4th -order Runge– Kutta numerical integration
techniques. The step size is fixed for most of the flight (𝑑𝑡 = 0.01 𝑠) but is made
smaller near the end of the flight (𝑑𝑡 = 0.0001 𝑠 when Range < 1000 ft) to accurately
capture the magnitude of the miss distance. The program is terminated when the
v𝑐 changes sign, because this means that the separation between the missile and
target is a min. At this time the missile-target separation is the miss distance.
We can develop a strategic ballistic missile-target engagement simulation by using
Earth-centered coordinate system 𝑥̈ 𝑡 = −𝐺𝑀 𝑥𝑡 /(𝑥𝑡2 + 𝑦𝑡2 )3/2 ; 𝑦̈ 𝑡 = −𝐺𝑀 𝑦𝑡 /(𝑥𝑡2 + 𝑦𝑡2 )3/2
where 𝐺𝑀 is the gravitational parameter. These differential equations are in an
inertial coordinate system. Therefore, they can be integrated directly to yield the
velocity and position of the target with respect to the Earth. The components of the
relative position between the missile and target can be expressed as

𝐫 = 𝐫𝑡 − 𝐫𝑚 and 𝐯𝑡𝑚 = 𝐫̇𝑡 − 𝐫̇𝑚 = 𝐯𝑡 − 𝐯𝑚

Application of the distance formula shows that the relative separation between the
0.5
missile and target can be found from 𝑟 = (𝑟𝑥2 + 𝑟𝑦2 ) . The closing velocity, which is
defined as the negative rate of change of separation between missile and target,
can be obtained by taking the negative derivative of the preceding expression,
yielding v𝑐 = −(𝑟𝑥 v𝑡𝑚𝑥 + 𝑟𝑦 v𝑡𝑚𝑦 )/𝑟. The LOS angle can be found by trigonometry as
𝜆 = tan−1 (𝑟𝑦 /𝑟𝑥 ). Therefore, the instantaneous value of the line-of-sight rate can be
found by taking the derivative of the preceding expression, using the quotient rule,
yielding 𝜆̇ = (𝑟𝑥 v𝑡𝑚𝑦 − 𝑟𝑦 v𝑡𝑚𝑥 )/𝑟 2. We now have sufficient information to guide a
strategic interceptor. The proportional navigation guidance command is
proportional to the LOS rate according to a𝑐 = 𝑁v𝑐 𝜆̇ = 𝑁v𝑐 (𝑟𝑥 v𝑡𝑚𝑦 − 𝑟𝑦 v𝑡𝑚𝑥 )/𝑟 2 this
guidance command is perpendicular to the LOS. And it can be seen that the
components of the guidance command in the Earth coordinate system can be
found by trigonometry and are given by a𝑚𝑥 = −a𝑐 sin 𝜆 & a𝑚𝑦 = a𝑐 cos 𝜆. Therefore,
the acceleration diff-equations describing the missile consist of two parts: the
gravitational term and the guidance command term. The components of the
missile diff-equations in Earth- coordinates are
−𝐺𝑀 𝑥𝑚 −𝐺𝑀 𝑦𝑚
𝑥̈ 𝑚 = + a𝑚𝑥 and 𝑦̈𝑚 = + a𝑚𝑦
(𝑥𝑚
2 2 )3/2
+ 𝑦𝑚 (𝑥𝑚
2 2 )3/2
+ 𝑦𝑚
Here is a MATLAB code for the simulation of an engagement between an impulsive
missile (i.e. strategic interceptor) and a ballistic target.

clear all, clc, dt=0.05; V=25; k=1; % you can use for eg V=150;
function dx=Missile(x,Vc,dL,L)
N=3.5; GM= 4.035e5; x1=x(1); x2=x(2); x3=x(3); x4=x(4);
dx1=x2; dx2=-GM*x1*(x1^2+x3^2)^(-1.5)-N*Vc*dL*sin(L);
dx3=x4; dx4=-GM*x3*(x1^2+x3^2)^(-1.5)+N*Vc*dL*cos(L);
dx=[dx1 dx2 dx3 dx4]';
end
function dxt=Traget(xt)
GM= 4.035e5; xt1=xt(1); xt2=xt(2); xt3=xt(3); xt4=xt(4);
dxt1=xt2; dxt2=-GM*xt1*(xt1^2+xt3^2)^(-1.5);
dxt3=xt4; dxt4=-GM*xt3*(xt1^2+xt3^2)^(-1.5);
dxt=[dxt1 dxt2 dxt3 dxt4]';
end
x(:,1)=[0 V*cos(-pi/3) 200 V*sin(-pi/3)]'; xt(:,1)=[500 -20 0 70]';
Rmtx= xt(1,1)-x(1,1); Rmty=xt(3,1)-x(3,1);
Vmtx=xt(2,1)-x(2,1); Vmty=xt(4,1)-x(4,1);
Rmt=(Rmtx^2+ Rmty^2)^0.5; Vc=-(Rmtx*Vmtx+Rmty*Vmty)/Rmt;
L= atan2(Rmty,Rmtx); dL=(Rmtx*Vmty-Rmty*Vmtx)/Rmt^2;
while Rmt>5
dx=Missile(x(:,k),Vc,dL,L); dxt=Traget(xt(:,k));
x(:,k+1)= x(:,k)+dx*dt; xt(:,k+1)= xt(:,k)+dxt*dt;
Rmtx= xt(1,k+1)-x(1,k+1); Rmty=xt(3,k+1)-x(3,k+1);
Vmtx= xt(2,k+1)-x(2,k+1); Vmty=xt(4,k+1)-x(4,k+1);
Rmt=(Rmtx^2+Rmty^2)^0.5; Vc=-(Rmtx*Vmtx+Rmty*Vmty)/Rmt;
L=atan2(Rmty,Rmtx); dL=(Rmtx*Vmty-Rmty*Vmtx)/Rmt^2; k=k+1;
end
figure; plot(x(1,:), x(3,:),'b','linewidth',1.5), grid on
hold on; plot(xt(1,:), xt(3,:),'r','linewidth',1.5), grid on
Thus far we have seen that
proportional navigation appears to be effective, but we do not know why. Although
it is possible to construct geometric arguments showing that it is very logical to
issue acceleration commands proportional to the line-of-sight rate (that is, zero
line-of-sight rate means we are on a collision triangle and therefore no further
commands are necessary), it is not obvious what is happening. The concept of zero
effort miss is not only useful in explaining proportional navigation but is also
useful in deriving and understanding more advanced guidance laws.

We can define the zero effort miss to be the distance the missile would miss the
target if the target continued along its present course and the missile made no
further corrective maneuvers. The zero effort miss can be expressed in terms of the
previously defined relative quantities as

𝑴𝑑 = 𝐫 + 𝐯𝑡𝑚 𝑡go = 𝐫 + 𝐯𝑡𝑚 (𝑡f − 𝑡); 𝑟 = ‖𝐫‖


where 𝐫 = 𝐫𝑡 − 𝐫𝑚 & 𝑡go = 𝑡f − 𝑡 = (v𝑐0 /𝑟) − 𝑡 is the 𝑴𝑑
time to go until intercept. Thus, we can see that 𝐯𝑡𝑚
in this case the zero effort miss is just a simple 𝑇
prediction (assuming constant velocities and zero 𝑀 𝐫
acceleration) of the future relative separation between missile and target. From
figure we can see that the component of the zero effort miss that is perpendicular
to the line of sight 𝑀⊥ can be found by trigonometry and is given by
cos(𝜆) sin(𝜆)
𝑀⊥ = [− sin(𝜆) cos(𝜆)]𝑴𝑑 with 𝑴oriented = [ ] 𝑴𝑑
− sin(𝜆) cos(𝜆)
From the other hand we have
𝑑𝜆 𝑡go 𝑀⊥
𝑀⊥ = vLOS 𝑡go = {𝑟 } 𝑡go ⟹ 𝑀⊥ = [𝑟𝑥 v𝑡𝑚𝑥 − 𝑟𝑦 v𝑡𝑚𝑦 ] and 𝜆̇ =
𝑑𝑡 𝑟 𝑟𝑡go
If we assume that the relative separation between missile and target and closing
velocity are approximately related to the time to go by: 𝑡go = 𝑟/vc then the
proportional navigation guidance command can be expressed in terms of the zero
effort miss perpendicular to the line sight as
𝑑𝜆 𝑟 𝑀⊥ 𝑁 𝑁 𝑁 𝐫
ac = 𝑁vc =𝑁 = 2 𝑀⊥ and 𝐚c = 𝑴⊥ = [𝑴𝑑 − (𝑴𝑑 . 𝐫) ]
𝑑𝑡 𝑡go 𝑟𝑡go 𝑡go 2
𝑡go 2
𝑡go ‖𝐫‖2
Thus, we can see that the proportional navigation acceleration command that is
perpendicular to the line of sight is not only proportional to the line-of-sight rate
and closing velocity but is also proportional to the zero effort miss and inversely
proportional to the square of time to go. Now let we summarize the algorithm in
Guidance law Update
𝐫 = 𝐫𝑡 − 𝐫𝑚 and 𝐯𝑡𝑚 = 𝐯𝑡 − 𝐯𝑚 𝐫𝑡 (𝑘 + 1) = 𝐫𝑡 (𝑘) + 𝐯𝑡 (𝑘)𝑑𝑡
𝑴𝑑 = 𝐫 + 𝐯𝑡𝑚 𝑡go |
𝐯𝑡 (𝑘 + 1) = 𝐯𝑡 (𝑘) + 𝐚𝑡 𝑑𝑡
𝑴∥ = (𝑴𝑑 . 𝐫)(𝐫/‖𝐫‖2 ) 𝐫𝑚 (𝑘 + 1) = 𝐫𝑚 (𝑘) + 𝐯𝑚 (𝑘)𝑑𝑡
𝑴⊥ = 𝑴𝑑 − (𝑴𝑑 . 𝐫)(𝐫/‖𝐫‖2 ) | 𝑁
2 𝐯𝑚 (𝑘 + 1) = 𝐯𝑚 (𝑘) + ( 2 ) 𝑴⊥ 𝑑𝑡
𝐚c = (𝑁/𝑡go )𝑴⊥ 𝑡go
%------------------------------------------------------------------
clear all, clc, figure; N=4; dt=0.01; t=0; w=3; At=6*9.81; k=1;
Rt(:,1)=[0 12192 3048]'; Rm(:,1)=[0 0 3048]'; %initial positions
Vm= [0 914.4 0]'; Vt=[-(At/w);- 304.8; 0]; %initial speed
tF=norm(Rt(:,1)-Rm(:,1))/norm(Vm-Vt); %predicted final time
%------------------------------------------------------------------
while t<tF-1e-5
tgo =(tF-t);
%------------------- Guidance law--------------------%
Vtx = -(At/w)*cos(w*t); Vty = - 304.8; Vtz = (At/w)*sin(w*t);
Vt=[Vtx; Vty; Vtz]; % Target velocity
Rtm = Rt(:,k)-Rm(:,k); % Relative position
Vtm = Vt-Vm; % Relative velocity
Md = Rtm + Vtm*tgo; % Miss-Distance
Mdn = Md-dot(Md,Rtm)*Rtm/(norm(Rtm))^2; % Normal Miss-Distance
%---------------------- Updates----------------------%
Rt(:,k+1)= Rt(:,k)+ Vt*dt; % Target position
Rm(:,k+1)= Rm(:,k)+ Vm*dt; % Missile position
Vm = Vm + (N*Mdn/tgo^2)*dt; % Missile velocity
t=t+dt; k=k+1;
end
%------------------------------------------------------------------
plot3(Rm(1,:),Rm(2,:),Rm(3,:),'b','linewidth',1.5), grid on, hold on
plot3(Rt(1,:),Rt(2,:),Rt(3,:),'r','linewidth',1.5), grid on
%------------------------------------------------------------------
Miss distance 𝐌𝑑 = 𝐫 − (𝐫. 𝐮𝑇/𝑀 )𝐮𝑇/𝑀 is often used as
a measure of missile system performance. In general, the smaller the miss
distance, the greater the probability of killing the target. Missile kill probability is a
function of many factors-including 𝐌𝑑 , fuze, warhead characteristics, and target
vulnerability. Acceptable miss distance (sufficiently small) is the first criterion of a
successful engagement because the fuze and warhead must be delivered relatively
close to the target in order to perform their functions. 𝐌𝑑 is usually defined as the
closest approach of some point on the missile-usually the missile center of mass-to
some point on the target-often the target center of mass. When miss distance is
defined as the closest approach of the missile center of mass to the target center of
mass, the closest approach occurs when the range vector 𝐫 reaches a minimum.

Relative target (𝐫. 𝐮𝑻/𝑴 )𝐮𝑻/𝑴


postion

𝐌𝑑 𝐌𝑑 = 𝐫 − (𝐫. 𝐮 𝑇/𝑀 )𝐮𝑇/𝑀


𝐫. 𝐮𝑇/𝑀
𝐫 𝑡𝑐𝑎 = 𝑡 −
Missile ‖𝐯𝑇/𝑀 ‖

The seeker tracking point is assumed to be the center of mass of the target and
although a first-order lag in the tracking rate is
introduced later, the small angular deviation of the seeker
boresight-axis vector from the line of sight to the tracking
point is not calculated. Also the displacement of the
physical position of the seeker from the missile center of 𝝈 𝐫
mass is considered negligible for this application.
Therefore, the seeker line-of-sight vector 𝝈 is assumed to
be identical with the range vector 𝐫: 𝝈 ≈ 𝐫 [m]. The angular
rate of the seeker line-of-sight vector is calculated by using the following relation

𝝎𝝈 = (𝝈 × 𝐯𝑇/𝑀 )/‖𝝈‖2 ≈ 𝝎gl [rad/s]

The angular rate of the seeker head lags the angular rate of the line-of-sight vector.
This lag is taken into account in calculating the guidance commands. This lag is
assumed to be represented by a first-order transfer function, and the seeker-head
rate is calculated by using
1
𝝎ach (𝑡) = 𝝎ach (𝑡 − ∆𝑡) exp(∆𝑡/𝜏1 ) + 𝝎𝝈 (1 − exp(∆𝑡/𝜏1 )) [rad/s] 𝝎𝝈 → → 𝝎ach
𝜏1 𝑠 − 1

There are assumed to be delays involved in processing the seeker-head angular


rate signal. The filtered seeker-head angular rate signal is given by
1
𝝎f (𝑡) = 𝝎f (𝑡 − ∆𝑡) exp(∆𝑡/𝜏2 ) + 𝝎ach (1 − exp(∆𝑡/𝜏2 )) [rad/s] 𝝎ach → → 𝝎f
𝜏2 𝑠 − 1

𝜏1 : seeker tracking loop constant and 𝜏2 : seeker signal processing time constant.
Radio frequency systems have the potential to measure the magnitude of the
closing velocity v𝑐 = ‖𝐯𝑐 ‖ = −(𝐫. 𝐯𝑇/𝑀 )/‖𝐫‖, i.e., magnitude of the velocity of the
missile relative to the target, and missiles with RF seekers sometimes implement
the "PN" proportional navigation by employing the closing speed v𝑐 = −𝑑𝑟/𝑑𝑡. A
practical implementation of PN for RF seekers is obtained by calculating the
commanded acceleration 𝐚c (𝑡) = 𝑁vc (𝝎f × 𝐮cl ) [m/s 2 ] where 𝑁: the navigation
constant (also known as navigation ratio, effective navigation ratio, and navigation
gain), a positive real number [dimensionless], 𝐮cl : unit vector in the direction of
centerline axis (from tile to nose of missile), the missile centerline is given by the
Euler angles: 𝐮cl = [cos 𝜃 cos 𝜓 cos 𝜃 sin 𝜓 − sin 𝜃]𝑇 , 𝝎f : The filtered seeker-head
angular rate signal [rad/s].

A guided missile engagement is a highly


dynamic process. The conditions that determine how close the missile comes to the
target are continuously changing, sometimes at a very high rate. A guidance
sensor measures one or more parameters of the path of the missile relative to the
target. A logical process is needed to determine the required flight path corrections
based on the sensor measurements. This logical process is called a guidance law.
The objective of a guidance law is to cause the missile to come as close as possible
to the target.

Command-to-line-of-sight guidance is similar to beam-rider guidance, in that both


forms attempt to keep the missile within a
guidance beam transmitted from the ground. As
Guideline
shown in figure, the vector 𝐞 represents the error
in missile position relative to the guidance beam at 𝐁
any given instant. This error is defined as the 𝐞
perpendicular distance from the missile to the
𝐏𝑩
centerline of the guidance beam. The missile
guidance commands generated by beam rider and 𝐏𝑴
command-to-line-of-sight systems are proportional
to the error vector 𝐞 and the rate of change of that
vector 𝐞̇ . The proportionality with 𝐞 causes the
missile to be steered toward the center of the
guidance beam; the proportionality with 𝐞̇ provides rate feedback, which causes
the missile flight path to maneuver smoothly onto the centerline of the guidance
beam without large overshoots.

A third parameter, the Coriolis acceleration 𝐴corio , may be included in the guidance
equation. This Coriolis acceleration results from the angular rotation of the
guidance beam. The Coriolis component of missile acceleration is required in order
to allow the missile to keep up with the rotating beam as the missile flies out along
the beam. In surface-to-air missile applications the angular rate of the guidance
beam is typically great enough to cause this parameter to be significant.
The error vector is 𝐞 = 𝑷𝐵 − 𝑷𝑀 where 𝐞 = vector of error in missile position relative
to the guideline, [m]. 𝑷𝐵 = position vector of a point on the guideline at the point of
intercept with the error vector 𝐞, [m]. 𝑷𝑀 = position vector of the missile, [m]. The
vector 𝑷𝐵 should be written in terms of the guideline and the missile position
vector. 𝑷𝐵 = (𝒖gl . 𝑷𝑀 )𝒖gl = (𝑷 𝑇 . 𝑷𝑀 /‖𝑷 𝑇 ‖2 )𝑷 𝑇 = 𝜅𝑷 𝑇 where: 𝒖gl = 𝑷 𝑇 /‖𝑷 𝑇 ‖ unit vector
that represents the direction of the guideline. The error rate vector 𝐞̇ is calculated
as 𝐞̇ = 𝐯𝐵𝑝𝑒𝑟 − 𝐯𝑀𝑝𝑒𝑟 = 𝝎gl × 𝑷𝐵 − (𝒖gl × 𝐯𝑀 ) × 𝒖gl , with 𝝎gl = (𝑷 𝑇 × 𝐯𝑇 )/‖𝑷 𝑇 ‖2 where
𝝎gl = is the angular rate vector of the guideline, [rad/s]. The Coriolis acceleration
term is calculated by 𝐴corio = Mag(𝝎gl × (𝒖gl . 𝐯𝑀 )𝒖gl ) where Mag[. ] = the magnitude
of the argument vector. Finally, using the terms calculated in 𝐞, 𝐞̇ and 𝐴corio the
commanded-lateral-acceleration vector, to guide the missile onto the centerline of
the guide beam, is

𝐚c = 𝑘1 𝐞 + 𝑘2 𝐞̇ + 𝑘3 𝐴corio 𝐮c 𝐮r = 𝐫/‖𝐫‖
𝐫 = 𝑷 𝑇 − 𝑷𝑀
= 𝑘1 𝐞 + 𝑘2 𝐞̇ + 𝑘3 𝐴corio (𝝎gl × 𝐮cl )/‖𝝎gl × 𝒖cl ‖
𝐮cl = [C(𝜃)C(ψ) C(𝜃)S(ψ) −S(θ)]𝑇
where 𝑘1 , 𝑘2 and 𝑘3 are proportionality constants (gains). 𝐮c = (𝝎gl × 𝐮cl )/‖𝝎gl × 𝐮cl ‖
unit vector in the direction of the component of 𝐞 that is perpendicular to the
missile centerline. The equation of 𝐚c represents the commanded-lateral-
acceleration vector that is fed to the control system to produce the convenient
deflection to minimize the error 𝐞. The choice of the proportionality constants 𝑘1 , 𝑘2
and 𝑘3 is usually computed by using optimization methods such as 𝐏𝐒𝐎 algorithm
to minimize the miss distance.

Proportional navigation is the most widely known


and used guidance law for short to medium range homing missiles, because of its
inherent simplicity and ease of implementation. Proportional navigation is so
robust, however, that acceptable miss distances can be achieved even against
targets that perform relatively severe evasive maneuvers if the missile response
time is short enough and if the missile is capable of sufficient acceleration in a
lateral maneuver.

The parallel navigation rule stats that: the


direction of the line-of-sight (LOS) is kept 𝑳𝑶𝑺𝟏
constant relative to inertial space, i.e., the
LOS is kept parallel to the initial LOS. The
𝑳𝑶𝑺𝟎
proportional navigation (PN) is the
guidance law which implements parallel 𝑳𝑶𝑺𝟐
T
navigation, but it kept the line-of-sight
rate to be zero rather than of constant
parallel navigation
direction. Therefore in PN the line of sight M
may changes direction but with zero rate.
PN dictates that the missile velocity vector should rotate at a rate proportional to
the rotation rate of the line of sight (LOS-rate), and in the same direction.
Simply stated, classical proportional navigation guidance is based on recognition
of the fact that if two bodies are closing on each other, they will eventually
intercept if the line of sight (𝑳𝑶𝑺) between
the two does not rotate relative to the
inertial space. It is based on the fact that Target
𝑳𝑶𝑺𝟎
two vehicles are on a collision course
when their direct line-of-sight does not
change direction as the range closes.
More specifically, the 𝑷𝑵 guidance law
seeks to null the 𝑳𝑶𝑺 rate against non-
maneuvering targets by making the 𝜆̇𝑳𝑶𝑺 = 0 or
interceptor missile heading proportional 𝜆𝑳𝑶𝑺 ≈ constant
to the 𝑳𝑶𝑺 rate. For instance, in flying a
proportional navigation course, the
missile attempts to null out any line-of- proportional navigation
Missile
sight rate that may be developing. The
missile does this by commanding wing deflections to the control surfaces.
Consequently, these deflections cause the missile to execute accelerations normal
to its instantaneous velocity vector. Thus, the missile commands wing deflections
to null out measured 𝑳𝑶𝑺 rate. The relation describing the normal commanded
acceleration is: 𝐚𝑐 = 𝑁v𝑐 (𝝎f × 𝐮cl ) ⟹ a𝑐 = 𝑁𝑉𝑐 𝜔f where a𝑐 : the normal commanded
acceleration [m/s2 ], 𝑁: the navigation constant, v𝑐 : the closing velocity [m/s],
𝜔f = 𝑑𝜆/𝑑𝑡 the LOS filtered rate measured by the missile seeker [rad/s].

Target

Target

𝛾 Seeker
lines of
Lines of sight
𝛾
sight

Missile
path

Fire unit

Beam_Rider Guidance Proportional Navigation PN

Almost all missiles that have seekers employ PN, which depends on the angular
tracking rate of the seeker head. Here we summarize the PN method
𝑑𝑟
𝝎gl = (𝐫 × 𝐯𝑇/𝑀 )/‖𝐫‖2 [rad/s] v𝑐 = − = −𝐫. 𝐯𝑇/𝑀 /𝑟 [m/s]
𝑑𝑡
𝝎ach (𝑡) = 𝝎ach (𝑡 − ∆𝑡)𝑒 (∆𝑡/𝜏1 )
+ 𝝎gl (1 − 𝑒 (∆𝑡/𝜏1 )
) [rad/s] | 𝐚 (𝑡) = 𝑁v (𝝎 × 𝐮 ) [m/s 2 ]
c c f cl
𝝎f (𝑡) = 𝝎f (𝑡 − ∆𝑡)𝑒 (∆𝑡/𝜏2 ) + 𝝎ach (1 − 𝑒 (∆𝑡/𝜏2 ) ) [rad/s] 𝐮cl = [c𝜃 c𝜓 c𝜃 s𝜓 −s𝜃 ]𝑇
𝜏1 : seeker tracking loop constant, and 𝜏2 : seeker signal processing time constant.
• PN dictates that the missile velocity vector should rotate at a rate proportional to
the rotation rate of the line of sight (LOS-rate), and in the same direction.

𝐚c (𝑡) = 𝑁(𝐯𝑐 × 𝝎LOS ) with 𝝎LOS = (𝐫 × 𝐯𝑇/𝑀 )/(𝐫. 𝐫)


Since LOS is not in general co-linear with the missile velocity vector, the applied
acceleration does not necessarily preserve the missile kinetic energy. In practice, in
the absence of engine throttling capability, this type of control may not be possible.

• Proportional navigation can also be achieved using acceleration normal to the


instantaneous velocity difference: 𝐚c (𝑡) = 𝑁(𝐯𝑇/𝑀 × 𝝎LOS ) with 𝝎LOS = (𝐫 × 𝐯𝑇/𝑀 )/(𝐫. 𝐫)
This acceleration depends explicitly on the velocity difference vector, which may be
difficult to obtain in practice.

• If acceleration normal to the instantaneous LOS is desired, then the following


expression is valid: 𝐚c (𝑡) = −𝑁‖𝐯𝑇/𝑀 ‖(𝐮𝑟 × 𝝎LOS ) with 𝐮𝑟 = 𝐫/‖𝐫‖.

• If energy conserving control is required (as is the case when only using control
surfaces), the following acceleration, which is orthogonal to the missile velocity,
may be used: 𝐚c (𝑡) = 𝑁‖𝐯𝑇/𝑀 ‖(𝐮𝑚 × 𝝎LOS ) with 𝐮𝑚 = 𝐯𝑚 /‖𝐯𝑚 ‖. A rather simple
hardware implementation of this guidance law can be found in AIM-9 Sidewinder
missiles. These missiles use a rapidly rotating parabolic mirror as a seeker.

clear all, clc, figure; N=4; dt=0.01; t=0; w=3; At=6*9.81;k=1; e=5;
Rt(:,1)=[0 12192 3048]'; Rm(:,1)=[0 0 3048]'; %initial positions
Vm= [0 914.4 0]'; Vt=[-(At/w);- 304.8; 0]; %initial speed
while e>2
%------------------- Guidance law--------------------%
Vtx = -(At/w)*cos(w*t); Vty = - 304.8; Vtz = (At/w)*sin(w*t);
Vt = [Vtx; Vty; Vtz]; % Target velocity
Rtm = Rt(:,k)-Rm(:,k); % Relative position
Vtm = Vt-Vm; % Relative velocity
Vc = -dot(Rtm,Vtm)/(norm(Rtm)); % Closing speed
W = cross(Rtm,Vtm)/(norm(Rtm))^2; % LOS-rate
ur = Rtm/norm(Rtm); um = Vtm/norm(Vtm); % Unit vectors
Ac = N*cross(Vtm,W); % Normal acceleration
% Ac = -N*norm(Vtm)*cross(ur,W);
% Ac = -N*Vc*cross(ur,W);
% Ac = N*norm(Vtm)*cross(um,W);
%---------------------- Updates----------------------%
Rt(:,k+1)= Rt(:,k)+ Vt*dt; % Target position
Rm(:,k+1)= Rm(:,k)+ Vm*dt; % Missile position
Vm = Vm + Ac*dt; % Missile velocity
t = t + dt; k = k+1; e = norm(Rtm);
end
plot3(Rm(1,:),Rm(2,:),Rm(3,:),'b','linewidth',1.5), grid on, hold on
plot3(Rt(1,:),Rt(2,:),Rt(3,:),'r','linewidth',1.5), grid on
This section we propose an improved
version of 3D pure pursuit navigation against a maneuvering target. Unlike
traditional 3D pure pursuit navigation, the guidance algorithm developed adapts
the direction, but maintains the magnitude of the commanded acceleration
proportional to the target acceleration. The validity and performance of the
proposed guidance algorithm are investigated through theoretical analysis and
numerical simulations. Let we start by computing the missile acceleration 𝐚𝑚 from
the well-known formula of the classical pure pursuit navigation
𝐫 𝑑v𝑚 𝐫 v𝑚 𝑑𝐫 𝑑𝑟 𝐫 v𝑚
𝐯𝑚 = v𝑚 ⟹ 𝐚𝑚 = + 2 [𝑟 ( ) − 𝐫 ( )] = a𝑚 − 2 [𝑟𝐯𝑐 + v𝑐 𝐫]
𝑟 𝑑𝑡 𝑟 𝑟 𝑑𝑡 𝑑𝑡 𝑟 𝑟
Therefore, the missile acceleration will be
a𝑚 v𝑚 v𝑐 v𝑚 𝑑𝐫
𝐚𝑚 = { − 2 } 𝐫 − 𝐯𝑐 ; with 𝐯𝑐 = − = closing velocity
𝑟 𝑟 𝑟 𝑑𝑡
Here, there is some complexity in the formula because in computing 𝐚𝑚 we need
first to evaluate a𝑚 which is not yet computed, so in order to overcome such
problem we assume that the a𝑚 is proportional to the target acceleration: a𝑚 = 𝑘a𝑡 .
a𝑡 v𝑚 v𝑐 v𝑚
𝐚𝑐 = {𝑘 − 2 }𝐫− 𝐯; with 𝑘 = proportionality constant
𝑟 𝑟 𝑟 𝑐
Missiles are controlled by self-contained automatic devices called accelerometers.
Accelerometers are inertial devices that measure accelerations. In missile control,
they measure the vertical, lateral, and longitudinal accelerations of the controlled
missile.
Guidance law Update
𝐫 = 𝐫𝑡 − 𝐫𝑚 and 𝐯𝑡𝑚 = 𝐯𝑡 − 𝐯𝑚
𝐯𝑐 = −𝑑𝐫/𝑑𝑡 = −𝐯𝑡𝑚 | 𝐫𝑡 (𝑘 + 1) = 𝐫𝑡 (𝑘) + 𝐯𝑡 (𝑘)𝑑𝑡
𝐫𝑚 (𝑘 + 1) = 𝐫𝑚 (𝑘) + 𝐯𝑚 (𝑘)𝑑𝑡
v𝑚 = ‖v𝑚 ‖; v𝑐 = ‖v𝑐 ‖; 𝑟 = ‖𝐫‖
| 𝐯𝑚 (𝑘 + 1) = 𝐯𝑚 (𝑘) + 𝐚𝑐 𝑑𝑡
a𝑡 v𝑚 v𝑐 v𝑚
𝐚𝑐 = {𝑘 − 2 } 𝐫 − 𝐯 𝑘 =𝑘+1
𝑟 𝑟 𝑟 𝑐
clear all, clc, N=20; dt=0.01; t=0; w=3; At=6*9.81; k=1; tF=15;
Rt(:,1)=[0 12192 3048]'; Rm(:,1)=[0 0 3048]'; %initial positions
Vm= [0 914.4 0]'; Vt=[-(At/w);- 304.8; 0]; %initial speed
while t<(tF-1e-5)
Vtx = -(At/w)*cos(w*t); Vty = - 304.8; Vtz = (At/w)*sin(w*t);
Vt = [Vtx; Vty; Vtz]; % Target velocity
Rtm = Rt(:,k)-Rm(:,k); % Relative position
Vtm = Vt-Vm; Vc = -Vtm; R=norm(Rtm);
Ac=((N*At*R-norm(Vm)*norm(Vc))*Rtm/R^2)-(norm(Vm)/R)*Vc;
Rt(:,k+1)= Rt(:,k)+ Vt*dt;
Rm(:,k+1)= Rm(:,k)+ Vm*dt;
Vm = Vm + Ac*dt;
t = t + dt; k = k+1;
end
figure; plot3(Rm(1,:),Rm(2,:),Rm(3,:),'b','linewidth',1.5), grid on
hold on, plot3(Rt(1,:),Rt(2,:),Rt(3,:),'r','linewidth',1.5)
Notes:
Parallel Navigation: The missile maintains a constant angle with the target's path,
effectively traveling in parallel to the target. The missile's course remains
unchanged unless the target changes its path. This method is simpler but less
effective for fast or maneuvering targets.

Proportional Navigation: The missile adjusts its flight path proportionally to the
rate of change of the line of sight (LOS) angle to the target. The missile constantly
steers towards the predicted future position of the target. It uses a gain factor
(navigation constant) that multiplies the LOS rate. More effective for intercepting
moving targets and compensating for target maneuvers. Continuously adjusts
trajectory based on the LOS rate. Proportional navigation is widely used in modern
missile guidance systems due to its robustness in various combat scenarios. It
aims to intercept the target by predicting its future position.

The proportional navigation (PN) is the guidance law which implements parallel
navigation, but it kept the line-of-sight rate to be zero rather than of constant
direction. PN can be seen as achieving a form of parallel navigation relative to the
moving target, as the missile's path continuously adjusts to maintain a direct
interception course. However, it fundamentally differs from maintaining a constant
direction or angle as in traditional parallel navigation.
• Introduction to Flight Simulation
SAM-Missile Flight • Simulation Synthesis
Simulation & Testing • Optimized Guidance Law by PSO
• Example of Simulation
(Verification and Validation) • Discussion and Conclusion

The users of a missile fright simulation


must have confidence that the simulation results are meaningful and that the
simulation output is representative of actual missile performance. It is essential
that the models of the missile system, subsystems, and physical environment have
a demonstrable correspondence with the system, subsystem, or environment being
modeled. This confidence is gained through the processes of verification and
validation. Verification ensures that the computer program operates correctly
according to the conceptual model of the missile system. Validation determines the
extent to which the simulation is an accurate representation of the real world.

Most if not all, flight simulations contain approximations and consequently are not
expected to be perfect representations of the actual missile system over all flight
conditions. One of the objectives of validation is to determine the flight conditions
for which the simulation does accurately represent the actual missile. Validation
is performed by comparing simulation output with right-test and laboratory data
obtained under similar flight conditions. Various methods are used to make these
comparisons; they range from visual comparison of plotted data overlays to
sophisticated statistical and spectral analyses. Missile flight simulations are often
developed progressively as the missile system is developed. As new and better data
on the actual system become available, the simulation model is updated, and the
validation of the model is extended to include the update.

The earlier chapters of this text described missile systems and methods used to
simulate the various missile subsystems. The purpose of this chapter is to show
how to synthesize a simulation by using the information provided in the earlier
chapters. An example of a relatively simple digital flight simulation of a generic
surface-to-air missile is used to illustrate the principals involved.

Equations and procedures for modeling the various


subsystems of missile flight have been presented in previous chapter. This chapter
employs an example to illustrate how the level of detail in a simulation is selected
to satisfy simulation objectives, and to show how to synthesize a complete flight
simulation by combining the subsystem models.

As discussed before, missile flight simulations are


developed to fulfill various objectives, and the details of the simulation depend
largely on those objectives. In the development of the example simulation,
objectives are selected that lead to a simulation sophisticated enough to illustrate
the principles but not so complicated that clarity is sacrificed.
The objectives of the example simulation are derived from the
following hypothetical scenario. A new surface- to-air missile system is to be
developed. The time is early in the development process. The missile configuration
is still in the conceptual phase, and a missile flight simulation is needed to
evaluate various design alternatives. Aerodynamic data are available for missiles
that are generally similar to the proposed configurations, but only limited wind
tunnel data are available for the specific configurations to be modeled. No flight-
test data are yet available. The autopilot and control systems have not been
defined in detail, but general transfer functions are available for the types of
systems that are likely to be developed for this missile. The seeker design
requirements have not been completed, but tentative seeker characteristics have
been estimated.

In the previous scenario, the overall objective of the proposed


simulation is that it be adequate to investigate the gross effects of different missile
design alternatives on missile system performance. Thus the simulation should be
constructed so that the missile and subsystem characteristics in it are easy to
change. The simulation must have at least five degrees of freedom in order for the
dynamic response characteristics of preliminary missile designs to be studied. It is
anticipated that the missile motion in roll about its longitudinal axis will be
sufficiently controlled so that the roll degree of freedom need not be simulated at
this point in the development. It is assumed that the missile will have cruciform
symmetry.

In a digital simulation the processing is done in discrete time


steps, the size of which must be carefully considered to ensure faithful
representation of the highest frequency components of the simulated missile
system. At any given time step the processing proceeds through each task and
calculates any changes that occur within that time increment. After completion of
all tasks appropriate to that time increment, the program goes to the next time
increment and repeats the cycle. This procedure is described in the following
diagram that shows the flow of processing from one task to the next.

Each block in the diagram represents a major function, or group of functions, or a


major logic process in the computer program (see the figure on the next page). The
direction of processing flow is indicated by arrows. One cycle through the flow
diagram represents an incremental time step.

Missile and target position and velocity vectors 𝐏𝑀 and 𝐏𝑇 are used to calculate the
relative position 𝐫 and velocity vectors 𝐯𝑇𝑀 with respect to the target. A test is
performed to determine whether the missile has reached its closest approach to
the target, which of course will not occur until the end of the engagement. If the
test shows that the closest approach has been reached, the program sequence is
diverted to a routine that calculates miss distance 𝐌𝑑 and the program ends.
Otherwise, the program continues into the guidance routine.
%-----------Initialization-----------% %------------UPDATING------------%
Give all the necessary initial values 𝐏𝑀 = 𝐏𝑀 + 𝐯𝑀 ∆𝑡; ℎ = −𝐏𝑀 (3);
%------------------------------------%
𝐴max ; 𝑡𝑚𝑖 ; 𝑃𝑑 ; 𝜏4 ; 𝑤𝑑 = 2𝜋/𝑃𝑑 ;
while 𝑡 ≤ 𝑡max
%------Atmosphere & Mach Number------% 𝐴 𝑇𝑐 = 𝐴max cos(𝑤𝑑 𝑡𝑚𝑖 ) ; 𝐸 = exp(−∆𝑡/𝜏4 );
ℎ1 = 0; 𝑃ref ; 𝑅 = 287.26; 𝑇 = 𝑇1 − 𝑎(ℎ − ℎ1 ); 𝐴Tach = 𝐴Tach 𝐸 + 𝐴 𝑇𝑐 [1 − 𝐸];
𝑃 = 𝑃ref exp[−g(ℎ − ℎ1 )/(𝑇𝑅)]; 𝜌 = 𝑃/(𝑅𝑇); 𝐰𝑡 = [0 0 𝐴Tach /‖𝐯𝑇 ‖];
v𝑠 = [𝛾𝑅𝑇]1/2; M𝑁 = ‖𝐯𝑀 ‖/v𝑠 ; 𝐀 𝑇 = 𝐰𝑡 × 𝐯𝑇 ; 𝐯𝑇 = 𝐯𝑇 + 𝐀 𝑇 ∆𝑡;
if (M𝑁 ∈ [… ]) % From table 𝐏𝑇 = 𝐏𝑇 + 𝐯𝑇 ∆𝑡 + 𝐀 𝑇 (∆𝑡 2 /2); 𝑡 = 𝑡 + ∆𝑡 ;
𝐶𝐷𝑜 = ⋯; 𝐶𝐿𝑎 = ⋯; 𝐶𝑚𝑎 = ⋯; if 𝑡 ≥ 𝑡𝑏𝑜 , 𝑚 = 𝑚𝑏𝑜 ; end
𝐶𝑚𝑠 = ⋯; 𝐶𝑚𝑛 = ⋯; 𝐾 = ⋯; ∆𝑥cmo = 𝑥cmo − 𝑥cmbo ; ∆𝑚𝑜 = 𝑚𝑜 − 𝑚𝑏𝑜 ;
end 𝑥cm = 𝑥cmo − ∆𝑥cmo [(𝑚𝑜 − 𝑚)/∆𝑚𝑜 ];
𝑄 = 0.5𝜌‖𝐯𝑀 ‖2 ; %dynamic pressure 𝐼 = 𝐼𝑜 − (𝐼𝑜 − 𝐼𝑏𝑜 )[(𝑚𝑜 − 𝑚)/∆𝑚𝑜 ];
%----Relative Position & Velocity----% 𝐮cl = [cos(𝜃) ; cos(𝜓) cos(𝜃) ; sin(𝜓) sin(−𝜃)];
𝐯𝑇𝑀 = 𝐯𝑇 − 𝐯𝑀 ; 𝐮𝑇𝑀 = 𝐯𝑇𝑀 /‖𝐯𝑇𝑀 ‖; 𝐮vm = 𝐯𝑀 /‖𝐯𝑀 ‖; 𝛼 = tan−1 (𝑤/𝑢);
𝑅prev = ‖𝐫‖; 𝐫 = 𝐏𝑇 − 𝐏𝑀 ; 𝑅next = ‖𝐫‖; 𝛽 = tan−1 (−𝑣/𝑢); 𝛼𝑡 = cos−1(𝐮vm . 𝐮cl );
if (𝑅prev − 𝑅next ) < 0 𝑅𝑇𝐷 = 180/𝜋; 𝑢1 = 𝛿𝑝 𝑅𝑇𝐷; 𝑢2 = 𝛿𝑦 𝑅𝑇𝐷;
𝑀𝑑 = ‖𝐫 − (𝐫. 𝐮𝑇𝑀 )𝐮𝑇𝑀 ‖; break end
end, 𝐮𝑟 = 𝐫/‖𝐫‖; V𝑐 = −(𝐮𝑟 . 𝐯𝑇𝑀 );
𝜏1 ; 𝜏2 ; 𝜏3 ; 𝑡gon ; 𝛿max ; Initialization
%---------Guidance & Control---------%
𝑁𝑟 ; 𝐺𝑛 ; 𝐰g = (𝐫 × 𝐯𝑇𝑀 )/‖𝐫‖2;
Atmosphere,
𝐰ach = 𝐰ach exp(−∆𝑡/𝜏1 ) + 𝐰g [1 − exp(−∆𝑡/𝜏1 )];
𝑀𝑁 , 𝑄
𝐰f = 𝐰f exp(−∆𝑡/𝜏2 ) + 𝐰ach [1 − exp(−∆𝑡/𝜏2 )];
𝐀c = 𝑁𝑟 ‖𝐯𝑀 ‖(𝐰f × 𝐮cl ); 𝐀cb = 𝐓𝑏𝑒 𝐀c % PN
𝛼𝑝 = −𝐺𝑛 𝐴𝑐𝑏 (3)/𝑄; 𝛼𝑦 = −𝐺𝑛 𝐴𝑐𝑏 (2)/𝑄; 𝐯𝑇𝑀 ; 𝐫
𝛼𝑝𝑎 = 𝛼𝑝𝑎 exp(−∆𝑡/𝜏3 ) + 𝛼𝑝 [1 − exp(−∆𝑡/𝜏3 )];
𝛼𝑦𝑎 = 𝛼𝑦𝑎 exp(−∆𝑡/𝜏3 ) + 𝛼𝑦 [1 − exp(−∆𝑡/𝜏3 )];
𝛿𝑝 = 𝛼𝑝𝑎 − 𝛼; 𝛿𝑦 = 𝛼𝑦𝑎 − 𝛽;
if 𝑡 < 𝑡gon , 𝛿𝑝 = 0; 𝛿𝑦 = 0; end Closest Yes
if abs(𝛿𝑝 ) > 𝛿max , 𝛿𝑝 = [sign(𝛿𝑝 )]𝛿max ; end Approach?
if abs(𝛿𝑦 ) > 𝛿max , 𝛿𝑦 = [sign(𝛿𝑦 )]𝛿max ; end
%-----------Aerodynamics-------------%
𝐶𝐿 = 𝐶𝐿𝑎 𝛼𝑡 ; 𝐶𝐷 = 𝐶𝐷0 + 𝐾𝐶𝐿 2 ; 𝐶𝑙 = 0; No
𝐿 = 𝑄𝑆𝐶𝐿 ; 𝐷 = 𝑄𝑆𝐶𝐷 ; 𝐴 = 𝐷 cos 𝛼𝑡 − 𝐿 sin 𝛼𝑡 ;
𝑁 = 𝐷 sin 𝛼𝑡 + 𝐿 cos 𝛼𝑡 ; ∆𝑥cm = 𝑥cm − 𝑥ref; Guidance
and Control
𝐅𝐴 = [−𝐴; (−𝑁𝑣)/√𝑣 2 + 𝑤 2 ; (−𝑁𝑤)/√𝑣 2 + 𝑤 2 ];
𝐌𝒅
𝐶𝑁𝑧 = 𝐹𝐴 (3)/(𝑄𝑆);𝐶𝑁𝑦 = 𝐹𝐴 (2)/(𝑄𝑆);
𝐶mref = 𝐶ma 𝛼 + 𝐶ms 𝛿𝑝 ; 𝐶nref = 𝐶ma 𝛽 + 𝐶ms 𝛿𝑦 ; Dynamics
𝐶m = 𝐶mref − 𝐶𝑁𝑧 (∆𝑥cm /𝑑) + 0.5𝑑𝑞(𝐶mn /‖𝐯𝑀 ‖); and Kinematics
𝐶n = 𝐶nref + 𝐶𝑁𝑦 (∆𝑥cm /𝑑) + 0.5𝑑𝑟(𝐶mn /‖𝐯𝑀 ‖);
𝐿𝑎 = 𝑄𝐶𝑙 𝑆𝑑; 𝑀𝑎 = 𝑄𝐶m 𝑆𝑑; 𝑁𝑎 = 𝑄𝐶n 𝑆𝑑; 𝐯𝑀 ; 𝐫𝑀
%-------Propulsion & GRAVITY---------%
if 𝑡 ≥ 5.6, Fref = 0; end,% Fref from table
F𝑝 = Fref + (𝑃ref − 𝑃𝑎 )𝐴𝑒 ; 𝐅𝑝 = [F𝑝 ; 0; 0]; Update
𝐹g𝑥 = [−𝑚g sin(𝜃)]; 𝐹g𝑦 = 𝑚g cos(𝜃) sin(𝜙);
𝐹g𝑧 = 𝑚g cos(𝜃) cos(𝜙); 𝐅g = [𝐹g𝑥 ; 𝐹g𝑦 ; 𝐹g𝑧 ];
%-----Solve by Rungr-Kutta Method----%
Get 𝐯𝑚 = [𝑢; 𝑣; 𝑤] and 𝐫𝑚 from the missile 𝑇 > 𝑇max Yes
No
𝑻𝑒𝑏 = Rotation Matrix; 𝑻𝑒𝑏 = 𝑻𝑇𝑏𝑒 ; 𝐯𝑀 = 𝑻𝑒𝑏 𝐯𝑚 ; or Crash? ? End
Optimization methods are widely used
in various fields. The task is to choose the best or a satisfactory one from amongst
the feasible solutions to an optimization problem. The process of using
optimization methods to solve a practical problem mainly involves these two steps.
First, formulate the optimization problem which involves determining the decision
variables, objective function and constraints, and possibly an analysis of the
optimization problem. Second, select an appropriate numerical method, solve the
optimization problem, test the optimal solution and make a decision accordingly.
Mathematically, an optimization problem may be summarized as follows: given a
criterion max( f(𝐱)) or min( f(𝐱)) where f(𝐱) is the objective function and 𝐱 is an n-
dimensional vector consisting of the decision variables.

Swarm intelligence refers to a class of algorithms that simulates


natural and artificial systems composed of many individuals that coordinate using
decentralized control and self-organization. The algorithm focuses on the collective
behaviors that result from the local interactions of the individuals with each other
and with the environment where these individuals stay. Some common examples
of systems involved in swarm intelligence are colonies of ants and termites, fish
schools, bird flock, animal herds. The particle swarm optimization (PSO) algorithm
falls into the category of SI algorithms and is a population-based optimization
technique originally developed by Kennedy and Eberhart in 1995. In PSO
algorithm, each agent is treated as a particle with infinitesimal volume with its
properties being described by the current position vector, its velocity vector and
the personal best position vector. Each agent knows the global best particle (𝐠 𝑏𝑒𝑠𝑡 )
between all the best particles (𝐱 𝑖𝑏𝑒𝑠𝑡 ).

1. Velocity Vector: denotes the increment of the current position. It is given for
each agent or (particle) by 𝐯𝑖𝑘+1 = 𝜎𝐯𝑖𝑘 + 𝛼𝐫1 × (𝐱𝑏𝑒𝑠𝑡
𝑖 − 𝐱𝑖𝑘 ) + 𝛽𝐫2 × (𝐠𝑏𝑒𝑠𝑡 − 𝐱𝑖𝑘 ) where

𝐯𝑖𝑘 = agent 𝑖 current velocity at the 𝑘 𝑡ℎ iteration,


𝜎 = inertia weight or (weighting function), takes a value between 0 and 1.
𝛼, 𝛽 = inertia weight factor, (learning parameters or acceleration constants)
𝐫1 , 𝐫2 = random vectors, and each entry takes a value between 0 and 1.
𝐱𝑖𝑘 = agent 𝑖 current position at the 𝑘 𝑡ℎ iteration,
𝐱 𝑖𝑏𝑒𝑠𝑡 = 𝑖 𝑡ℎ agent’s best position, and 𝐠 𝑏𝑒𝑠𝑡 = group’s best position.
2. Inertia Weight: is given as follows 𝜎(𝑘) = 𝜎max − (𝜎max − 𝜎min ) × (𝑘/𝑘max ). In the
simplest case, the inertia function 𝜎(𝑘) can be taken as a constant, typically
𝜎(𝑘) ∈ [0.5 0.9]. This is equivalent to introducing a virtual mass to stabilize the
motion of particles, and thus the algorithm is expected to converge more quickly.
3. Position Vector: is modified according to the equation 𝐱𝑖𝑘+1 = 𝐱𝑖𝑘 + 𝐯𝑖𝑘+1 where 𝐱𝑖𝑘
and 𝐱𝑖𝑘+1 are agent current and modified positions, 𝐯𝑖𝑘+1 = agent modified velocity.

The initial locations of all particles should be distributed relatively uniformly so


that they can sample over most regions, which is especially important for
multimodal problems. The initial velocity of a particle can be set to zero, (i.e. 𝐯𝑖𝑘=0 ).
%----maximum value of the objective function f(𝑥, 𝑦) = 3 sin(𝑥) + 𝑒 𝑦 , −4 ≤ 𝑥, 𝑦 ≤ 4----
clear all, clc,[X,Y]=meshgrid(-4:.5:4,-4:.5:4); Z=3*sin(X)+exp(Y); surf(X,Y,Z);
wmax=0.9; wmin=0.4; c1=1.49; c2=1.49; itermax=50; xmin=[-2 -2]; xmax=[2 2];
n=20; m=2; v=zeros(m,n); rand('state',0);
%-------------------------------------------------------------------------------
for i=1:n
for j=1:m, x(j,i)=xmin(j)+rand*(xmax(j)-xmin(j)); end
fun_marge(i)=3*sin(x(1,i))+exp(x(2,i));
end
xbest=x; fbest=fun_marge;
fgbest=min(fun_marge);
gbest=x(:,find(fun_marge==fgbest));
%-------------------------------------------------------------------------------
for iter=1:itermax
w=wmax-(wmax-wmin)*iter/itermax;
for i=1:n
v(:,i)=w*v(:,i)+c1*rand*(xbest(:,i)-x(:,i))+c2*rand*(gbest-x(:,i));
x(:,i)=x(:,i)+v(:,i);
for jj=1:m
if x(jj,i)>xmax(jj), x(jj,i)=xmax(jj); end
if x(jj,i)<xmin(jj), x(jj,i)=xmin(jj); end
end
fun_marge(i)=3*sin(x(1,i))+exp(x(2,i));
if fun_marge(i)<fbest(i), xbest(:,i)=x(:,i); fbest(i)=fun_marge(i); end
if fun_marge(i)<fgbest, gbest=x(:,i); fgbest=fun_marge(i); end
end
result(iter)=fgbest;
end
%-------------------------------------------------------------------------------
figure, fprintf(' the optimal value is %3.4f\n', gbest)
fprintf(' the minimum value of func is %3.4f\n', fgbest)

In the simulation results we’ve used this optimization technique and we’ve defined
the miss distance to be the function subject to the minimization. A MATLAB code
of the PSO algorithm applied to the missile is provided at the end.
A simulation is based on mathematical models of the
missile, target and environment, and these mathematical models consist of
equations that describe physical laws and logical sequences. The missile model
includes factors such as missile mass, thrust aerodynamics, guidance and control,
and the equations necessary to calculate the missile attitude and flight path. The
target model is often less detailed but includes sufficient data and equations to
determine the target flight path. The model of the environment contains, at a
minimum, the atmospheric characteristics and gravity.

For purposes of illustration it is assumed that a particular missile configuration is


to be investigated. The missile configuration to be studied is controlled by torque-
balanced canard control surfaces, and the canards and stabilizing tail fins are
arranged in a cruciform configuration. The description of the missile required for
the simulation model is given by

𝑚0 = 85.0 missile mass at launch, [kg]. 𝑚𝑏0 = 57.0 missile mass at burnout, [kg] .
𝐼0 = moment of inertia about x, y and z axes at launch, [kg. m2 ] 0.7 0 0
𝐼𝑏0 = moment of inertia about x, y and z axes at burnout, [kg. m2 ] [𝐼0 ] = [ 0 61 0 ]
𝑥𝑐𝑚0 = 1.55 distance frome nose to center of mass at launch, [m] 0 0 61
𝑥𝑐𝑚𝑏0 = 1.35 distance frome nose to center of mass at burnout, [m] 0.45 0 0
d = 0.127 aerodynamic refefence length (Missile’s diameter), [m] [𝐼𝑏0 ] = [ 0 47 0]
𝑙 = 1.6 fuselage length (length from tile to center of mass), [m] 0 0 47

Time (𝑡) 0 0.01 0.04 0.05 0.08 0.10 0.20 0.30 0.60 1.00 1.50
Trust (ref) 0 450 17800 23100 21300 20000 18200 17000 15000 13800 13300

Time (𝑡) 2.50 3.50 3.80 4.00 4.10 4.30 4.50 4.70 4.90 5.20 5.60 100
Trust (ref) 13800 14700 14300 12900 11000 7000 4500 2900 1500 650 0 0

𝑡𝑏0 = 5.6 time of burnout, [s] S = 0.0127 missile aerodynamic ref area, [m2 ]
𝑝ref = 101314 ref_ambient pressure, [Pa] d = 0.127 aerodynamic refefence length, [m]
𝐴𝑒 = 0.011 exit area of rocket nozzle, [m2 ] 𝑥ref = 1.35 distance from missile noze to
𝐼𝑠𝑝 = 2224 specific impulse, [N. s/kg] reference moment station, [m]
𝑉𝑠 = 343 sound speed in elastic medium [m/s] a = lapse rate 6.5 × 10−3 [K/m]
R = 287.26 gas constant, [N. m]/(kg. K) 𝜌0 = 1.2230 air density at sea level [kg/m3 ]
𝑅𝑒 = 6378 earth radius at equator, [km] 𝑇1 = 288.1667 → 290 [K] temp_at sea level

Mach Number 𝐌𝐍 , [dimensionless]


Coefficient
0.0 0.8 1.14 1.75 2.5 3.5
𝐶𝐷0 0.8 0.8 1.2 1.15 1.05 0.94
𝐶𝐿𝛼 38.0 39.0 56.0 55.0 40.0 33.0
𝐶𝑚𝛼 −160.0 −170.0 −185.0 −235.0 −190.0 −150.0
𝐶𝑚𝛿 180.0 250.0 230.0 130.0 80.0 45.0
𝐶𝑚𝑞 + 𝐶𝑚𝛼̇ −6.000 −13.000 −16.000 −13.500 −10.000 −6.000
𝑘 0.0255 0.0305 0.0361 0.0441 0.0540 0.0665
Initialization: The initial missile speed is assumed to apply the instant the missile
leaves the launcher. The direction of missile velocity will be calculated in the fire-
control routine. 𝐏𝑀 (𝑡 = 0) = [𝑟𝑥 (0) 𝑟𝑦 (0) 𝑟𝑧 (0)] = (0, 0, 0) missile initial position [m]
V𝑀 = ‖𝐯⃗𝑴 ‖ = 30, magnitude of missile initial velocity (in earth coordinate system), [m/s]
𝑡𝑚𝑎𝑥 = 60, maximum time of flight, [s]. ∆𝑡 = 5 × 10−3 integration time step, [s]
g 0 = 9.80665 acceleration due to gravity [m/s 2 ]. 𝜋 = 3.141592654 pi, [dimensionless]
𝐷𝑇𝑅 = 𝜋/180 factor converting radiance to degrees, [dimensionless], 𝑅𝑇𝐷 = 180/𝜋
[𝑝, 𝑞, 𝑟] = 𝟎, [rad/s] (deg/s) 𝛼 = 𝛽 = 𝛼𝑡 = 0, [rad] (deg)
𝐏𝑇 (𝑡 = 0) = [4 1 −3] target initial position [km]
𝐯𝑇 (𝑡 = 0) = [−250 0.00 0.00] target initial velocity, [m/s]
The initial missile inertial velocity vector 𝐯⃗𝑴 expressed in earth coordinates is
transformed into body coordinates using the earth-to-body reference frame
transformation matrix given in subroutine
𝑢 Cθ Cψ Cθ Sψ −Sθ 𝑣𝑥
(𝐯⃗𝑴 )𝑏 = [𝑻𝑏/𝑒 ](𝐯⃗𝑴 )𝑒 ⇔ [ 𝑣 ] = [Sϕ Sθ Cψ − Cϕ Sψ Sϕ Sθ Sψ + Cϕ Cψ Sϕ Cθ ] [𝑣𝑦 ] , [m/s]
𝑤 b Cϕ Sθ Cψ + Sϕ Sψ Cϕ Sθ Sψ − Sϕ Cψ Cϕ Cθ 𝑣𝑧 e
T
[𝑻𝑏/𝑒 ] = [𝑻𝑒/𝑏 ] transformation matrix (earth to body coordinates), [dimensionless]
𝑢, 𝑣, 𝑤: components of missile inertial velocity 𝐯⃗𝑴 in body coordinate system. [m/s]
𝐯⃗𝑀 = missile inertial velocity vector in earth coordinate system, [m/s. ]
At start when 𝑡 = 0 we put 𝑚(𝑡 = 0) = 𝑚0 , 𝐼(𝑡 = 0) = 𝐼0 , 𝑥𝑐𝑚 (𝑡 = 0) = 𝑥𝑐𝑚0
𝑟𝑥 ⃗𝑷⃗ 𝑇 − ⃗𝑷
⃗𝑀
−1 𝑥
𝑟 (0) −𝑟𝑧 (0) 𝐫
𝜓 = tan { −1
} , 𝜙 = 0, 𝜃 = tan { ⃗ (𝑡) = (𝑟𝑦 ) =
}, 𝒓 =
𝑟𝑦 (0) √𝑟𝑥 (0)2 + 𝑟𝑦 (0)2 𝑟𝑧 ‖𝐫‖ ‖𝑷⃗⃗ 𝑇 − ⃗𝑷
⃗ 𝑀‖

For this example of simulation, it is assumed that the missile altitude at the
launch position is at sea level; therefore, missile altitude above sea level, for use in
the atmosphere tables, is given by ℎ = −𝑷𝑀 (3); 3rd compoenet of 𝑷𝑀 in earth fram, [m]
%===============MISSILE FLIGHT SIMULATION=================%
clear all, close all, clc, format
velocity = []; altitude = []; temperature = []; gravity = [];
pressure = []; airdensity = []; massv = []; inertia = [];
centerg = []; pousse = []; attackangle = []; sideslipe = [];
machnumber = []; soundspeed = []; pitch = []; yaw = [];
roll = []; totalattack = []; lift = []; drag = [];
Pm = []; Pt=[]; u1=[]; u2=[];
y1=[]; y2=[]; y3=[]; y4=[]; y5=[]; y6=[]; y7=[]; y8=[]; y9=[];
Ts = 0.8; % sampling period [sec]
mo = 85.0; % missile mass at launch [Kg]
mbo = 57.0; % missile mass at burnout [Kg]
Io=[0.7 0 0;0 61 0;0 0 61]; % moment of inertia at launch [Kg.m^2]
Ibo=[0.45 0 0;0 47 0;0 0 47]; % moment of inertia burnout [Kg.m^2]
xcmo = 1.55; % distance from nose to center of mass at launch [m]
xcmbo = 1.35; % distance from nose to center of mass at burnout [m]
d0 = 0.127; % aerodynamic reference length(missile's diameter [m]
l = 1.6; % length from missile tile to center of mass [m]
S = 0.0127; % missile aerodynamic reference area [m^2]
xref = 1.35; % distance from nose to reference moment station [m]
rhoo = 1.223; % air density at sea level [Kg/m^3]
T1 = 288.1667; % temperature at sea level [K]
a = 0.0065; % lapse rate [K/m]
Rg = 287.26; % gas constant [N.m/(Kg.k)]
Re = 6378000; % earth radius at equator [m]
go = 9.80665; % acceleration due to gravity [m/s^2]
pref = 101314; % reference ambient pressure [pa]
gama = 1.4 ; % ratio of specific heat
Ae = 0.011; % exit area of rocket nozzle [m^2]
Isp = 2224; % specific impulse [N.s/Kg]
PM = [0 0 0]; % missile's initial position
PT = [4000 1000 -3000] ; % target's initial position
VT = [-250 20 50]; % target initial velocity [m/s]
VMn = 30; % magnitude of missile initial velocity
R = PT-PM; % range vector from missile to target
ur = R/norm(R); % unit range vector
ucl = ur; % unit centerline vector at launch
p = 0; q = 0; r = 0; % initial angular velocities [°/s]
VM=VMn*ucl; tita=atan(-ucl(3)/sqrt((ucl(1))^2+(ucl(2))^2));
phi = 0; psi = atan(ucl(2)/ucl(1)); alpha = 0; beta = 0; alphat = 0;
Tbe = [cos(psi)*cos(tita) cos(tita)*sin(psi) -sin(tita);...
-sin(psi) cos(psi) 0;...
cos(psi)*sin(tita) sin(psi)*sin(tita) cos(tita)];
Teb=Tbe'; Vm=Tbe*VM'; u=Vm(1); v=Vm(2); w=Vm(3); VM=VM';
tbo = 5.6; % time of burnout [s]
tmax = 60; % maximum time of flight [s]
Deltat=0.005; % integration time step [s]
t=0; % initial time
xcm=1.55; % distance from nose to center of mass at launch [m]
I=[0.7 0 0;0 61 0;0 0 61]; % moment of inertia at launch [Kg.m^2]
rho = 1.223; %air density at sea level [Kg/m^3]
pres = 101314 ; mass = 85.0; % missile mass at launch [Kg]
grav = 9.80665; % acceleration due to gravity [m/s^2]
tem = 288.1667; % temperature at sea level [K]
alphapa=0; alphaya=0; wach=[0 0 0]; wf=[0 0 0]; ATach=0;
yo=[p q r u v w phi tita psi]'; ppp=[0 0 0]; alt=0; thetaa=0; psia=0;
titaint=0; kk=1;[fref,mass] = trust1;
%------------------------------------------------------------------------%
while t <= tmax
%---------------- ATMOSPHERE----------------%
alt0=0; tem=T1-a*(alt-alt0); pres=pref*exp((-go./(tem*Rg))*(alt-alt0));
rho=pres/(Rg*tem);
%---------------- MACH NUMBER----------------%
vs = sqrt(gama*Rg*tem); Mn = norm(VM)/vs;
if(Mn>=0 & Mn<0.8)
CDo=0.8; CLa=38; Cma=-160; Cms=180; Cmn=-6000; K=0.0255;
end
if(Mn>=0.8 & Mn<1.14)
CDo=0.8; CLa=39; Cma=-170; Cms=250; Cmn=-13000; K=0.0305;
end
if(Mn>=1.14 & Mn<1.75)
CDo=1.2; CLa=56; Cma=-185; Cms=230; Cmn=-16000; K=0.0361;
end
if(Mn>=1.75 & Mn<2.5)
CDo=1.15; CLa=55; Cma=-235; Cms=130; Cmn=-13500; K= 0.0441;
end
if(Mn>=2.5 & Mn<3.5)
CDo=1.05; CLa=40; Cma=-190; Cms=80; Cmn=-10000; K=0.0540;
end
if(Mn>=3.5)
CDo=0.94; CLa=33; Cma=-150; Cms=45; Cmn=-6000; K=0.0665;
end
Q = 0.5*rho*(norm(VM)).^2; % dynamic pressure
%----------RELATIVE POSITION AND VELOCITY----------%
Vtm = VT-VM'; %relative velocity
utm=Vtm./norm(Vtm); Rprev=norm(R); R=PT-PM; Rnext=norm(R);
if (Rprev-Rnext)<0
Md=R-dot(R,utm)*utm; Md=norm(Md); % miss distance
tca = t - dot(R,utm)/norm(Vtm) % time of closest approaches
disp('*******************************************')
disp('*------ THE TARGET HAS BEEN DESTROYED ----*')
disp('*--------------- GAME OVER ---------------*')
disp('*******************************************')
break
end
ur = R/norm(R); Vc = -dot(ur,Vtm); % unit range and closing speed
tau1=0.01; tau2=0.01; tau3=0.04; tgon=0.5; sigmamax=0.3491;
%------------PROPORTIONAL NAVIGATION-------------%
% lamdam = 40; NR=4; gam =R; Gn=254.1; % usa = gam/(norm(gam));
% lamda = acos(dot(usa,ucl));
% wg = (cross(gam',Vtm)/norm(gam)^2);
% wach = wach*exp(-Deltat/tau1) + wg*(1-exp(-Deltat/tau1));
% wf = wf*exp(-Deltat/tau2) + wach*(1-exp(-Deltat/tau2));
% Gs = NR*norm(VM); Ac = Gs*(cross(wf,ucl)); Acb = Tbe*Ac';
%------------BEAM RIDER AND COMMAND TO LINE OF SIGHT-------------%
k1 = 4.15475; k2 = -0.041; k3 = 0.1151; Gn = 250;
wgl = cross(PT,VT)/(norm(PT))^2; ugl = PT./norm(PT);
PB=(dot(ugl,PM))*ugl; VMp=cross(cross(ugl,VM),ugl); VBp=cross(wgl,PB);
e = PB-PM; ed = VBp - VMp; uc = cross(wgl,ucl)./norm(cross(wgl,ucl));
Acc = norm(cross(wgl,(dot(VM,ugl).*ugl)));
Ac = k1*e + k2*ed + k3*Acc*uc; Acb = Tbe*Ac';
%------------------------------------------------------------------------%
%--------------------GUIDANCE AND CONTROL---------------------%
alphap=-Gn*Acb(3)/Q; alphay=Gn*Acb(2)/Q;
alphapa=alphapa*exp(-Deltat/tau3)+alphap*(1-exp(-Deltat/tau3));
alphaya=alphaya*exp(-Deltat/tau3)+alphay*(1-exp(-Deltat/tau3));
sigmap = alphapa - alpha; sigmay = alphaya - beta;
if t<tgon, sigmap=0; sigmay=0; end
if abs(sigmap)>sigmamax, sigmap=sign(sigmap)*sigmamax; end
if abs(sigmay)>sigmamax, sigmay=sign(sigmay)*sigmamax; end

%------------------AERODYNAMICS-------------------%
CL = CLa*alphat; CD = CDo + K*CL^2;
L = Q*S*CL; D = Q*S*CD; A = D*cos(alphat)-L*sin(alphat);
N = D*sin(alphat)+L*cos(alphat);
Fa = [-A N*(-v/sqrt(v^2 + w^2)) N*(-w/sqrt(v^2 + w^2))];
Cnz = Fa(3)/(Q*S); Cny = Fa(2)/(Q*S);
Cmref = Cma*alpha + Cms*sigmap; Cnref = Cma*beta + Cms*sigmay;
Cm = Cmref - Cnz*((xcm-xref)/d0)+ (d0/(2*norm(VM)))*(Cmn)*q;
Cn = Cnref + Cny*((xcm-xref)/d0)+ (d0/(2*norm(VM)))*(Cmn)*r;
La = Q*S*d0*0; Ma = Q*S*d0*Cm; Na = Q*S*d0*Cn;
%%-------PROPULSION-------%%
if t >=5.6, fref(kk)=0; end, fp=fref(kk)+(pref-pres)*Ae; Fp=[fp 0 0];
%%-------GRAVITY-------%%
Fgx=-mass(kk)*grav*sin(tita);
Fgy= mass(kk)*grav*cos(tita)*sin(phi);
Fgz= mass(kk)*grav*cos(tita)*cos(phi);
Fg=[Fgx Fgy Fgz]; pousse = [pousse;vs];
%%-----------------RUNGR-KUTTA Method-------------%%
H = Deltat; HH = H/2; H6 = H/6;
V=yo; dydx=derivs(mass(kk),I,La,Ma,Na,Fa,Fg,Fp,V); yt = yo + HH*dydx';
V=yt ; dyt=derivs(mass(kk),I,La,Ma,Na,Fa,Fg,Fp,V); yt = yo + HH*dyt';
V=yt ; dym=derivs(mass(kk),I,La,Ma,Na,Fa,Fg,Fp,V); yt = yo + H*dym';
dym = dyt + dym ; V=yt ; dyt=derivs(mass(kk),I,La,Ma,Na,Fa,Fg,Fp,V);
yout = yo + H6*(dydx' + dyt' + 2*dym');
%----------------------------------------------------------------%
d=yout; p=d(1); q=d(2); r=d(3); u=d(4); v=d(5); w=d(6); phi=d(7);
tita = d(8);psi = d(9); yo = [p q r u v w phi tita psi]';
y1 = [y1;p];y2 = [y2;q];y3 = [y3;r];y4 = [y4;u];y5 = [y5;v];
y6 = [y6;w];y7 = [y7;phi];y8 = [y8;tita];y9 = [y9;psi];
Tbe = [cos(psi)*cos(tita) cos(tita)*sin(psi) -sin(tita);...
-sin(psi) cos(psi) 0;...
cos(psi)*sin(tita) sin(psi)*sin(tita) cos(tita)];
Teb = Tbe'; Vm = [u;v;w]; VM = Teb*Vm;
%--------------- MISSILE POSITION---------------%
PM = PM + VM'.*(Deltat); % missile position
% PT = PT + VT.*(Deltat); % target linear position
alt = -PM(3); altitude = [altitude; alt]; % missile altitude
%---------------TARGET POSITION---------------%
ng = L/mass(kk)*grav ; % load factor
Amax = grav.*sqrt(4); tmi=1; Pd=120; tau4=0.01;
% A_Tc = Amax*sign(cos(2*pi*tmi/Pd));
ATc = Amax*cos(2*pi*tmi/Pd);
ATach = ATach*(exp(-Deltat/tau4)) + ATc*(1-exp(-Deltat/tau4));
wt = [0 0 ATc/norm(VT)] ;
AT = cross(wt,VT); VT = VT + AT*Deltat;
PT = PT + VT.*(Deltat) + AT.*Deltat^2/2; % target curved position
%------------------------------------------------------------------------%
%---------------------UPDATING--------------------%
t = t + Deltat; kk = kk +1;
if t>=tbo, mass(kk) = 56.618; end
massv = [massv;mass(kk)];
xcm = xcmo - (xcmo-xcmbo)*((mo-mass(kk))/(mo-mbo));
centerg = [centerg;xcm];
I = Io - (Io-Ibo)*((mo-mass(kk))/(mo-mbo));
inertia = [inertia;(diag(I))'];
ucl = [cos(tita)*cos(psi) cos(tita)*sin(psi) sin(-tita)];
uvm = VM/norm(VM); alpha = atan(w/u); beta = atan(-v/u);
alphat = acos(dot(uvm,ucl)); % alphat = acos(u/norm(VM));
Pm = [Pm;PM]; Pt = [Pt; PT]; RTD = 180/pi;
u1 = [u1;sigmap*RTD]; u2 = [u2;sigmay*RTD];
temperature = [temperature;tem]; gravity = [gravity;grav];
airdensity = [airdensity;rho]; pressure = [pressure;pres];
attackangle = [attackangle;alpha*RTD];
totalattack = [totalattack;alphat*RTD];
sideslipe = [sideslipe;beta*RTD];
machnumber = [machnumber;Mn]; soundspeed=[soundspeed;vs];
pitch = [pitch;Ma]; yaw=[yaw;Na]; lift=[lift;L]; drag=[drag;D];
end
%------------------------******-------------------------%
tt =0.005*[0:length(u1)-1]'; figure, xm=Pm(:,1); ym=Pm(:,2); zm=Pm(:,3);
xt = Pt(:,1); yt = Pt(:,2); zt = Pt(:,3);
plot3(xm,ym,-zm,'r'), hold on, grid on, plot3(xt,yt,-zt,'b')
xlabel('x axis'), ylabel('y axis'), zlabel('z axis'),
title('FLIGHT SIMULATION'), legend('Missile','Target')
%------------------------------------------------------------------------%
function [a,b] = trust1()
x=[0 .01 .04 .05 .08 .1 .2 .3 .6 1.0 1.5 2.5 3.5 3.8 4.0 4.1...
4.3 4.5 4.7 4.9 5.2 5.6];
y=[0 450 17800 23100 21300 20000 18200 17000 15000 13800 13300 ...
13800 14700 14300 12900 11000 7000 4500 2900 1500 650 0];
xq=(0:0.005:5.6); vq=interpn(x,y,xq,'pchip');
t=0:0.005:5.6; n=length(t); pp=spline(x,y); ye=ppval(pp,x); time=0.005;
m0=85;
for i=1:n
m(:,i)= m0-(integral(@(x)ppval(pp,x),0,time))/2224; time = time+0.005;
end, a=vq ; b=m;
end
%------------------------------------------------------------------------%
function h = derivs(mass,I,La,Ma,Na,Fa,Fg,Fp,y)
%----------------angular velocities-------------------------
yout(1) = (La - y(2)*y(3)*(I(3,3) - I(2,2)))/I(1,1);
yout(2) = (Ma - y(1)*y(3)*(I(1,1) - I(3,3)))/I(2,2);
yout(3) = (Na - y(1)*y(2)*(I(2,2) - I(1,1)))/I(3,3);
%----------------linear velocities--------------------------
yout(4) = -(y(2)*y(6)-y(3)*y(5))+(Fa(1)+Fg(1)+Fp(1))/mass;
yout(5) = -(y(3)*y(4)-y(1)*y(6))+(Fa(2)+Fg(2))/mass;
yout(6) = -(y(1)*y(5)-y(2)*y(4))+(Fa(3)+Fg(3))/mass;
%---------------Euler angles (accelerations)----------------
yout(7) = y(1)+(y(2)*sin(y(7))+y(3)*cos(y(7)))*tan(y(8));
yout(8) = y(2)*cos(y(7))-y(3)*sin(y(7));
yout(9) = (y(2)*sin(y(7))+y(3)*cos(y(7)))/cos(y(8));
h = yout;
end
%------------------------------------------------------------------------%
In this section, we present the simulation
results obtained when the missile fly, without acting on the fins (no deflection is
made), in order to visualize missile parameters variations and its trajectory.

The left figure shows that the missile attends its peak of thrust at the first instants
of its flight because the missile needs a maximum of energy to fly so far as possible
with highest velocity. At this point the thrust begins to decrease under the
influence of the resistance of the air versus the missile. The right figure illustrates
the change in mass of the missile during its flight. As the figure shows, we can
decompose the missile mass evolution into two phases:
Phase 01: begins at the instant of launch to 5.6 s, in this interval, we can see that
the mass of the corresponding missile decrease from 85 kg to 57 kg, however, this
decreasing is referred to the consumption of the propellant during the flight.
Phase 02: it is the last time of simulation, where the mass become constant (i.e. no
thrust, no fuel consumption), the missile fly only with its proper mass.
This first three figures show the evolution of inertia moment on x, y and z axes
respectively, this variation is related linearly with changing in mass by a
mathematical equation described in the missile modeling chapter. The last figure
shows that as the missile mass changes, the position of the center of mass also
changes linearly with it.

The two figures show the evolution of the atmospheric parameters during the
missile flight. As the missile’s altitude increase the atmospheric parameters
decrease on the first half time of flight until the missile gets its highest level of
altitude. However, these parameters increase in the second half of flight because
the missile’s altitude decreases. So these parameters are inversely proportional
with missile’s altitude.
The first figure presents the evolution of Mach number during the flight as
function of time, it is clearly seen that the Mach number increases rapidly in the
first 5 s until its highest value (Mach = 2.3 at t= 5.6 s), this refers to the enormous
thrust of missile at the beginning of its flight. After this point the missile Mach
number begins to decrease because of the termination of thrust.

The second figure illustrates the missile launch trajectory during its flight without
acting on the deflection surfaces (the missile in this case is like a projectile).

Now, we’ll present different simulation results when applying various guidance
laws to direct the missile toward the target. We will restore the capacity of each
method to keep track the target in each maneuvers, straight and curved path with
shortest time and lowest miss-distance. In the coming results some parameters
such as: thrust, mass, moment of inertia, center of mass, atmospheric parameters
(temperature, pressure, air-density and speed of sound), gravity and Mach number
are not presented because they aren't affected by the use of guidance laws.

In this simulation the target is


flying at an altitude of 3 km and a speed of 250 m/s, and the target flight path is
straight and offset laterally 1 km from the missile launch site. At the instant of
missile launch the target is inbound at a downrange distance of 4 km from the
launch site. The time of simulation is limited at 8 s, because we need 7.6 s at most
to destruct the target. The target position vector at a given time is calculated by
using: 𝐏𝑇 = 𝐏𝑇0 + 𝐯𝑇 𝑡. In the light of this straight flight we will see the

Proportional Navigation results


Beam Rider (Command to Line of Sight) results
Proportional Derivative Guidance results

And each method of them will be treated separately.


In this section we let missile to engage the target in
straight flight path by using proportional navigation guidance law.

The above figure illustrates the variation in deflection surfaces during the flight. As
we see, guidance is not initiated until a short time called "time to go on 𝑡gon " after
launch in order to permit the missile to gain enough speed so that it can be
controlled. After this point the autopilot bases the missile maneuver commands on
the achieved seeker-head angular rate vector, and the control system responds to
autopilot commands by deflecting the control surfaces. In the early time of flight
the fins are deflected violently because the seeker seeks to track the line of sight,
once he got it the deflections will be smooth as possible.
The angle of attack and sideslip angle histories that result from the moments
applied to the missile are shown in Figure 6.21. During the half second before
guidance is initiated, the angle of attack begins to increase slightly because gravity
causes the missile flight path to deviate downward from the direction the missile is
initially pointed as it leaves the launcher. The restoring moment, caused by this
small angle of attack, rotates the missile downward to point into its relative wind;
this reduces the angle of attack essentially to zero by the time guidance is initiated
and the missile begins to track the desired angles and tries to remain there.

The pitch and yaw rotational moments on the missile caused by the combination
of fin deflections, the restoring moment from the resulting angle of attack, and the
damping effect of the missile angular rate are shown figures. When the control fins
are initially deflected a large moment is generated and the missile rotates and
overshoots the trim angle of attack. A restoring moment is generated to rotate the
missile back toward the trim condition; this results an oscillation. The damping
moment causes oscillations to diminish until trim conditions are achieved.
This figure shows the missile and target trajectories during the flight, and their
interception in 3D. As we see in this figure the launcher is aimed directly at the
target at the time of launch, the proportional navigation guidance causes the
missile to turn in a direction to lead the target as is required to strike a moving
target. This missile maneuver is initiated when guidance is turned on (0.5 s). At
this early time in the flight, the missile speed is slow, which causes the amount of
lead, and, therefore, the amount of the maneuver to be overestimated. As the
missile gains speed, the missile flight path is corrected until intercept with target
at 𝑡 = 7.4587 s with miss-distance = 0.004 m as we see in the simulation result.

In this section we let missile


to engage the target in straight flight path by using beam rider guidance law.
The above figures show the variation of different parameters during the missile
engagement when beam rider or CLOS guidance used. The dynamics seems to be
the same, in other words there isn't a remarkable difference between the three
guidance laws in terms of the dynamics.

The first two figures illustrates the variation of the pitch and yaw rotational
moments caused by the difference of pressure applied on the control surfaces, the
figure show some chattering which can be minimized by the damping effect.

From the flight path figure we can see that the center-line axis of missile is
directed toward the target at each instant during the flight until the interception
point (or at closet approach), where the miss-distance = 0.4655 m at 𝑡 = 7.5197 s.
This is the geometry of the Beam rider guidance where the missile’s centerline is
always pointing toward the target.
In this section we let the missile to
engage the target in straight flight path by using proportional derivative navigation
guidance law.

%------------ Proportional Derivative COMMAND -------------%


K1 = [10 5.5839 8.4240; −10 − 8.0385 7.2738];
K 2 = [−2.6266 9.8147 6.3610; 4.1722 − 4.1955 − 8.6924];
𝐞 = 𝐏T − 𝐏M ; 𝐞̇ = 𝐯T − 𝐯M ; 𝜎 = K1 tanh(𝐞) + K 2 tanh(𝐞̇ );
𝜎𝑝 = 𝜎(1); 𝜎𝑦 = 𝜎(2); 𝐀 c = [0 𝜎𝑝 𝜎𝑦 ]; 𝐀 cb = 𝐓𝑏𝑒 𝐀𝑇c ;
Discussion: Through the above simulation figures, we have illustrated the effect of
this command on the evolution of each parameter during the time of simulation.
This method is easily implemented and the desired tracking performance can be
obtained by suitably selecting the controller gains using PSO algorithm. The use of
the tanh(. ) or equivalently sign(. ) function in the lateral acceleration equation
served to minimize oscillations. Last figure shows the two paths of missile and
target, and their interception point at t = 7.4508 s with miss-distance = 1.0492 m.

In this case the target is flying at an altitude of


3 km. The target flight path is curved and offset laterally 1 km from the missile
launch site. At the instant of missile launch the target is inbound at a downrange
distance of 4 km from the launch site. The weaving flight path can be modeled by
calculating the maneuver acceleration as a function of time by using:

𝐀 T = 𝛚T × 𝐯T (𝑡 − ∆𝑡)
𝐯T (𝑡) = 𝐯T (𝑡 − ∆𝑡) + 𝐀 T ∆𝑡
𝐏T (𝑡) = 𝐏T (𝑡 − ∆𝑡) + 𝐯T (𝑡 − ∆𝑡)∆𝑡 + 𝐀 T (∆𝑡)2 /2

where
𝐯T (𝑡) = the current target speed.
𝐯T (𝑡 − ∆𝑡) = the previous target speed.
𝐀 T = the maneuver acceleration of the target.
𝐏T = the target position vector.
𝛚T = the angular rate vector of the target flight path, rad/s
In this section we let the missile to engage
the target in weaving flight path by using proportional navigation guidance law.
Discussion: The above figures show the variation of different parameters during
the missile engagement when PN used. The 3D plot in last figure shows clearly the
geometry of PN which is the constant bearing angle kept between the LOS and the
missile centerline.
In this section we let the missile to
engage the target in weaving flight path by using beam rider guidance law.
Discussion: The above figures show the variation of different parameters during
the missile engagement when LOS guidance applied. The fist two figures shows
that the control signal was aggressive and oscillating; this may cause instability
and uncertainty for the missile.
In this section we let the missile to
engage the target in weaving flight path by using PD based guidance law.
Discussion: The above figures from show the variation of different parameters
during the missile engagement when PD guidance applied. The remarkable thing
which make this method differs from the others is that the yaw deflection surface
deflects in the negative sense after initiating guidance, which make the side-slipe
angle oscillates in the negative sense as shown in figure, that in turn affect the
variation of the yaw moment.

The last figure shows the missile and the target trajectories and their point of
intercept in 3 dimensions. It can be seen that the missile flies approximately in
straight line toward the target
Now let we do the comparison between
the three results in terms of time of closest approach and the miss-distance.

4.4.1 Target in straight flight


Proportional CLOS guidance PD based
navigation (Beam Rider) guidance
Miss-distance 0.0040 [m] 0.4655 [m] 1.0492 [m]
Time of closest approach 7.4587 [s] 7.5197 [s] 7.4508 [s]

4.4.2 Target in weaving flight


Proportional CLOS guidance PD based
navigation (Beam Rider) guidance
Miss-distance 0.1612 [m] 0.4597 [m] 0.4799 [m]
Time of closest approach 7.4275 [s] 7.5070 [s] 7.3853 [s]

Discussion: According to the above comparison tables, which illustrate the missile
performances in both straight and curved path, we observe that

Proportional navigation command shows good tracking, and records the lowest
miss-distance in shortest time, thus it is the most well-known used guidance law.
The command to line of sight has also good results in miss-distance, even it took
more time than the other two methods. The PD command marked acceptable
results due to the utilization of optimized gains. The tables also show that the PN
scored the lowest miss-distance, while the PD scored the shortest time of closest
approach.

Another criterion that is not mentioned in this study, which is the energy content
of the control signal (i.e. commanded accelerations) showed that the PD based
guidance law generates very high control signals. These signals must be limited by
limiters and conditions on the commanded accelerations to make this method
more realistic. In general, the dynamics still the same but the performances
depend on the applied guidance law.

Conclusion: In this chapter, we’ve presented a comparative study between three


guidance laws, namely the proportional navigation, the beam-rider or LOS
command guidance and PD based guidance law, when applied to the dynamics of
a generic surface to air missile. The three synthesized guidance laws proportional
navigation, the beam-rider or LOS command guidance and PD based guidance law,
were compared in terms of their performance to achieve the control objective and
their characteristics as well. Overall, we conclude from this study that the PN
outperforms the other methods in terms of miss-distance, while PD guidance is
better in terms of time of closest approach.
General Conclusion • Accuracy of Nonlinear Models
• Guidance of Military Vehicles
(with perspectives & suggestions) • Perspectives & Suggestions

Nonlinear Model can accurately reflect the real behavior of the missile because it
takes into account nonlinear dynamics. It is very useful in designing control
systems that need high precision, such as controlling the missile at large angles or
high speeds. It requires complex calculations and can be difficult to verify and
analyze, which increases the complexity of the system.

Proportional navigation and pure pursuit are mainly used in missile guidance
systems and military air vehicles to intercept moving targets. The PN method
depends on adjusting the missile's trajectory so that the rotation rate of the line of
sight between the missile and the target is proportional to the approach speed.
The pure pursuit guidance is a course in which the missile velocity vector is
always directed toward the instantaneous target position. The criterion for the
pure pursuit course is that the missile always heads directly toward the present
target position. Therefore, if the missile rate of turn equals the line-of-sight rate of
turn, the flight path becomes a pursuit path. Navigation by line-of-sight method
involves directing the missile (or drone) directly towards the target by maintaining
a constant line of sight between the missile (or drone) and the target. Control
commands are sent continuously to correct the missile's path.

Proportional navigation, line of sight and pure pursuit technology are not used
directly in civil aircraft. Civil aircraft mainly rely on traditional control systems
such as autopilot and flight management systems (FMS) that use different
technologies to ensure safe and stable flight. Some concepts similar to proportional
navigation may be used in some advanced aircraft guidance and landing systems,
but they are not based on the same principles used in military missiles. Controls
the civil aircraft based on pre-determined parameters such as altitude, speed, and
trajectory. Flight Management System (FMS), precisely manages the aircraft's
course, fuel control and navigation using advanced flight data.

For military applications, proportional, command to line of sight and pure pursuit
navigations are very effective at tracking and intercepting moving targets, such as
missiles or other aircraft. All of these technologies are used in an integrated
manner to achieve specific goals, such as intercepting enemy aircraft or hitting
ground targets accurately. For military applications a fighter aircrafts and drones
are used in surveillance and reconnaissance missions where all those vehicles
need to track a specific target. Moreover, fighter aircrafts uses all the above
described guidance laws in guided weapon systems to achieve precision strikes.

PN, line-of-sight control, and pure tracking techniques are used in missiles and in
fixed-wing UAVs based on the mission type and specific requirements. These
techniques are most effective in military applications, pure tracking is more widely
used in civilian applications.
The work conducted in this text can be extended in different directions. For
instance, one may reconsider the modeling of disturbances and introduce the
concept of noise and filtering even for the target or missile, one also may
reconsider a target making random evasive maneuvers. Moreover, for such type of
developed models more sophisticated guidance laws could be applied and tested,
for example higher order sliding mode, adaptive controllers, and augmented or
extended PN.

Also, we propose the use of neural networks and artificial intelligence to focus on
enhancing precision, adaptability, and robustness in various operational
scenarios. Integration of machine learning algorithms and artificial intelligence can
provide a sophisticated adaptive guidance laws. These systems can learn from data
and adapt to changing conditions in real-time, improving accuracy against
dynamic targets and in complex environments.

Utilization of multiple sensors (such as radar, infrared, and GPS) improves target
tracking and guidance. Fusion algorithms integrate data from these sensors to
enhance situational awareness and maintain guidance accuracy under diverse
conditions (e.g., cluttered environments, jamming).

The application of advanced nonlinear control techniques such as model predictive


control (MPC) and data based control algorithm provide robust performance
against uncertainties and disturbances while optimizing missile trajectories for
better interception capabilities.

We suggest a development of cooperative guidance strategies where multiple


missiles or platforms can coordinate their actions to achieve mission objectives
collaboratively. Networked guidance enables distributed decision-making and
enhances overall system effectiveness.

One can explore a deep reinforcement learning (DRL) algorithms for autonomous
decision-making in missile guidance systems. DRL enables missiles to learn
optimal actions through trial and error, improving adaptive response capabilities
in complex and uncertain environments.

These trends highlight a shift towards more intelligent, adaptable, and


autonomous missile guidance systems capable of operating in complex and
challenging environments while maintaining high precision and reliability.
Continued research and development in these areas are driving the evolution of
missile guidance technologies to meet future defense and security needs.
Appendix
1. MATLAB Code for the PSO Algorithm
clear all, close all, clc, format
wmax=0.9; wmin=0.8; c1=0.49; c2=0.49; rand('state',0);
n=20; itermax=30; xmin=[-2 -2 -0.2]; xmax=[-5 5 0.5]; m=3; v=zeros(m,n);

for i=1:n,
for j=1:m, x(j,i)=xmin(j)+rand*(xmax(j)-xmin(j)); end
e(i) = PSO_MIMO_MISSILE(x(1,i),x(2,i),x(3,i)) ;
fun_marge(i) = e(i);
end
xbest = x;
fbest = fun_marge;
fgbest = min(fun_marge);
gbest = x(:,find(fun_marge==fgbest));

for iter=1:itermax
w = wmax-(wmax-wmin)*iter/itermax;
for i=1:n
v(:,i) = w*v(:,i)+c1*rand*(xbest(:,i)-x(:,i))+c2*rand*(gbest-x(:,i));
x(:,i) = x(:,i)+v(:,i);
for jj = 1:m
if x(jj,i)>xmax(jj), x(jj,i)=xmax(jj); end
if x(jj,i)<xmin(jj), x(jj,i)=xmin(jj); end
end
e(i) = PSO_MIMO_MISSILE(x(1,i),x(2,i),x(3,i)) ; fun_marge(i) = e(i);
if fun_marge(i)<fbest(i), xbest(:,i)=x(:,i); fbest(i)=fun_marge(i); end
if fun_marge(i)<fgbest, gbest=x(:,i); fgbest=fun_marge(i); end
end
result(iter)=fgbest;
end

fprintf(' the optimal value is %3.4f\n', gbest)


fprintf(' the minimum value of func is %3.4f\n', fgbest)
e = PSO_MIMO_MISSILE(gbest(1),gbest(2),gbest(3));
plot([1:itermax], result,'--r','linewidth',1.5)
xlabel('Iteration'), ylabel('Function'), grid on

2. Numerical Solution of Differential Equations: Differential equations of first


order can be solved using variety of mathematical tools. But for solving the
equations using different initial conditions and real time inputs, we need a
computer generated approximate solution. This is where numerical integration
techniques come handy. Any normal system of differential equations can be
written as a first-order normal system, which in vector notation has the form

𝑑𝐱/𝑑𝑡 = 𝑭(𝑡, 𝐱)
Where
𝑭(𝑡, 𝐱) = vector o function of 𝑡 and 𝐱
𝑡 = independent variable (time)
𝐱 = vector of dependent variables.
The general solution of this differential equation is given by 𝐱 = 𝐟(𝑡)
The Runge_Kutta method and its variations are very popular simulations. The
method provides good accuracy, is simple to program, requires minimum storage,
and is stable under most circumstances with integration intervals of reasonable
size. The most common form of the method is based on the summation of four
terms; consequently, it is referred to as the fourth-order Runge_Kutta method. In the
fourth-order Runge_Kutta method, the most frequently selected arbitrary constants
lead to a set of difference equations of the form.

𝐱 𝑛+1 = 𝐱 𝑛 + [𝑲1 + 2𝑲2 + 2𝑲3 + 𝑲4 ]
6
Where
𝑲1 = 𝑭(𝑡𝑛 , 𝐱 𝑛 ),
𝑲2 = 𝑭(𝑡𝑛 + ℎ/2, 𝐱 𝑛 + ℎ𝑲1 /2),
𝑲3 = 𝑭(𝑡𝑛 + ℎ/2, 𝐱 𝑛 + ℎ𝑲2 /2),
𝑲4 = 𝑭(𝑡𝑛 + ℎ, 𝐱 𝑛 + ℎ𝑲3 ),
ℎ = integration step size
𝐱 𝑛 = vector of dependent variables at beginning of step 𝑛
𝐱 𝑛+1 =vector of dependent variables at the next step 𝑛 + 1

3. Coordinate system: A number of different coordinate systems may be used in a


given missile flight simulation. Coordinate systems are characterized by the
positions of their origins, their angular orientations, and their motions relative to
inertial space or relative to other specified systems. A given vector can be described
by its coordinates in any of the coordinate systems. If the coordinates of a vector
are given in one reference frame, the coordinates of that vector in any other
reference frame can be determined if the position and orientation of one reference
frame relative to the other is known.

3.1 Earth Coordinate System (𝑥𝐸 , 𝑦𝐸 , 𝑧𝐸 ) In a flat-earth simulation the earth is


usually assumed to be fixed in space, i.e., neither translating nor rotating. In this
case, absolute accelerations can be measured with respect to any coordinate
system fixed to the earth. Such a system is called an earth coordinate system and
is commonly used as a basis for measuring accelerations, velocities, and positions
of a missile, target, and decoys.

3.2 Body Coordinate System (𝑥𝐵 , 𝑦𝐵 , 𝑧𝐵 ) The body coordinate system is fixed to the
missile and aligned with the principal axes of the missile. Thus the system is
particularly useful for calculations of angular rates because the equations of
motion contain no terms involving the products of the moments of inertia and the
moments of inertia about the reference frame axes are independent of missile
attitude.

3.3 Wind Coordinate System (𝑥𝑊 , 𝑦𝑊 , 𝑧𝑊 ) The movement of undisturbed air relative
to the missile (relative wind) is tangent to the missile flight path. The wind
coordinate system is viewed as being aligned with the relative wind to simplify the
calculation of aerodynamic forces and moments. By definition, the aerodynamic
drag and lift vectors are aligned with wind system axes.
Bibliography

[1] U.S. Army Missile Command "Military Handbook: Missile Flight Simulation, Part
One, Surface-To-Air Missiles ", MIL-HDBK-1211(MI) 17 July 1995.

[2] Bureau of Naval Personnel "Principles of Guided Messiles and Nuclear Weapons",
Officer Candidate School, NavPcrs 10784-A, University of Texas, 1966.

[3] Bureau of Naval Weapons "Weapons Systems Fundamentals: Analysis of


Weapons", NAVWEPS OP 3000 (Volume 2), Washington, 1964.

[4] Air Training Command. United States, Air Force. "Fundamentals of Guided
Missiles", Aero Publishers, Inc. Los Angeles, California (1960)

[5] Naval Training Command "Gunner’s Mate Missiles 1 and C", United States
Government Printing Office Washington. 1973.

[6] Naval Training Support Command "Gunner’s Mate Missiles 3 and 2", United
States Government Printing Office Washington. 1972.

[7] Charles T. Myers "Guided Missiles Operations, Design and Theory", Mcgraw-Hill
Book Company, Inc, New York Toronto London 1958

[8] N. F. Twining, "Fighter Weapons", United States, Department Of The Air Force
Washington 25, D.C. May 1956.

[9] P. Garnell, D.J. East "Guided Weapon Control Systems", Pergamon Press Inc.
Maxwell House, Royal Military College of Science, New York 10523, U.S.A, 1977.

[10] Arthur S. Locke "Principles of Guided Missile Design", U.S. Naval Air
Development Center, D. Van Nostrand Company, Inc (1955)

[11] Joseph J. Jerger "Systems preliminary design", U.S. Naval Air Development
Center, D. Van Nostrand Company, Inc. (1960)

[12] N. A. Shneydor "Missile Guidance and Pursuit: Kinematics, Dynamics and


Control", Israel Institute of Technology, Coll House, Westergate, Chichester, West
Sussex, PO20 6QL England, 1998.

[13] Rafael Yanushevsky "Modern missile guidance", Kiev Polytechnic Institute,


Ukraine. Taylor, Francis Group LLC, 2019.
[14] Paul Zarchan "Tactical and strategic missile guidance", American Institute of
Astronautics and Aeronautics, Inc., Washington, DC, 1997

[15] Paul Zarchan, "Fundamentals of Kalman Filtering a Practical Approach".


American Institute of Aeronautics, Inc., Washington, DC, 2009

[16] S.R. Mohan "Fundamentals of Guided Missiles", Defense Research and


Development Organization, New Delhi, 2016
[17] John H. Blakelock "Automatic Control of Aircraft and Missiles", Wiley-Inter-
science publication, 1991.

[18] S. N. Balakrishnan, A. Tsourdos, B. A. White "Advances in Missile Guidance,


Control, and Estimation", Taylor, Francis Group, 2013.

[19] R. Balakrishnan "Guided Weapons System Design", Director, DESIDOC,


Metcalf House, Delhi-110 054, 1998.

[20] George M. Siouris "Missile guidance and control systems" Springer-Verlag New
York , Inc , 2004.

[21] Jack N. Nielsen "Missile Aerodynamics", McGRAW-Hill Book Company Inc.


New York Toronto London 1960

[22] Qi Zaikang, Lin Defu "Design of Guidance and Control Systems for Tactical
Missiles", Taylor, Francis Group, LLC, Beijing Institute of Technology, 2020.

[23] Warren J. Boord, John B. Hoffmanv "Air and Missile Defense Systems
Engineering", Taylor, Francis Group, LLC, 2016.

[24] Ashish Tewari "Atmospheric and Space Flight Dynamics Modeling and
Simulation with MATLAB and Simulink", Department of Aerospace Engineering.
Indian Institute of Technology, Kanpur India, 2007

[25] Swee, John CS. "Missile terminal guidance and control against evasive targets"
Naval Postgraduate School, Monterey, California 2000.

[26] D Harkin "Missile Guidance by Jet Impulse" Technical Report, Naval Research
Lab Washington Dc (Defense Technical Information Center) 1947.

[27] Gregory George Voulgarakis "Missile Miss Distance Reduction", Naval


Postgraduate School, Monterey, California 1982

[28] Ching-Fang Lin "Modern Navigation, Guidance, and Control Processing",


Prentice Hall Inc. 1991.

[29] Lukenbill, Francis C. "A target missile engagement scenario using classical
proportional navigation" Monterey, California: Naval Postgraduate School 1990

[30] Filippo Neri "Introduction to Electronic Defense Systems" -SciTech Publishing


(Artech House Radar Library) 2006

[31] Andrea De Martino "Introduction to Modern EW Systems" -SciTech Publishing


(Artech House Radar Library) 2012

[32] David Adamy "A second course in electronic warfare" -Artech House (2004)

[33] Clarence J. Lang "Fire Control Systems—General", Engineering Design


Handbook, Headquarters, U.S. Army Materiel Command 1968

[34] Bureau of Ordnance "Basic Fire Control Mechanisms", Navy Department


Bureau of Ordnance Washington 25. DC. 1944.
[35] Bureau of Naval Personnel "Fire Control Fundamentals", NAVPERS 91900,
Navy Department, Bureau of Ordnance, Washington 1953

[36] Bureau of Naval Personnel "Submarine Torpedo Fire Control Manual", Navy
Department, Bureau of Ordnance, Washington 1952

[37] Bureau of Naval Personnel "40 MM Antiaircraft Gun", TM 9-1252, Navy


Department, Bureau of Ordnance, Washington 1951

[38] Bureau of Naval Personnel "20 MM Antiaircraft Gun", OP 911, Navy


Department, Bureau of Ordnance, Washington 1943

[39] Stoje Deskovski "Homing Guidance and SA Missile Systems" Macedonian 2004

[40] K. L. Nielsen, J. F. Heyda "The Mathematical Theory of Airborne Fire Control"


Department of the Navy Bureau of Ordnance Washington 1951

[41] S. Krishnan "Pursuit Course In Aerial Gunnery" Defense Science Organization,


Ministry of Defense, New Delhi. 1956

[42] G. H. Handelman and W. R. Heller, “Aerodynamic Lead Pursuit Curves for


Overhead Attacks,” Applied Mathematics Panel Report 106.2 R. 1949

[43] Arthur Frederick Bernhart, “Curves of Pursuit”, University of Oklahoma,


Scripta Math 1957

[44] Arthur Frederick Bernhart, “Curves of General Pursuit”, University of


Oklahoma, Scripta Math 1959

[45] F. V. Morley, “Curves of Pursuit”, Taylor & Francis, Ltd. on behalf of the
Mathematical Association of America 1921

[46] C.C. Pucket “The Mathematical Gazette the Curve Of Pursuit” The Mathematical
Gazette, Volume 37, pp. 256-260 (1953)

[47] Patricia A. Hawley and Ross A. Blauwkamp “Six-Degree-of-Freedom Digital


Simulations for Missile Guidance, Navigation, and Control” Johns Hopkins Apl
Technical Digest, Volume 29, Number 1 (2010)

[48] Patricia A. “Overview of Missile Flight Control Systems” Johns Hopkins Apl
Technical Digest, Volume 29, Number 1 (2010)

[49] Neil F. Palumbo “Guest Editor’s Introduction: Homing Missile Guidance and
Control” Johns Hopkins Apl Technical Digest, Volume 29, Number 1 (2010)

[50] Neil F. Palumbo “Guidance Filter Fundamentals” Johns Hopkins Apl Technical
Digest, Volume 29, Number 1 (2010)

[51] Neil F. Palumbo “Modern Homing Missile Guidance Theory and Techniques”
Johns Hopkins Apl Technical Digest, Volume 29, Number 1 (2010)

[52] Neil F. Palumbo “Basic Principles of Homing Guidance” Johns Hopkins Apl
Technical Digest, Volume 29, Number 1 (2010)
[53] Brian L. Stevens “Aircraft control and simulation _ dynamics, controls design,
and autonomous systems” John Wiley & Sons, Inc 2016

[54] Donald McLean “Automatic-flight-control-systems” Prentice Hall International


(UK) Ltd 1990

[55] Nguyen X. Vinh “Flight Mechanics of High-Performance Aircraft” Cambridge


University Press 1993

[56] M.V. Cook “Flight Dynamics Principles” Published by Elsevier Ltd 2007

[57] Harold T. Davis “Introduction to Nonlinear Differential and Integral Equations”


United States Atomic Energy Commission 1960

[58] Jun Sun, Choi-Hong Lai, Xiao-Jun Wu "Particle Swarm Optimization Classical
and Quantum Perspectives". Taylor and Francis Group, LLC, 2012

[59] Slami Saadi. "Introduction a l’optimisation meeta-heuristique", Dar Kifaya, Bab-


Ezzouar, Algeria, 2019.

[60] Shinar, J "Three Dimensional Optimal Pursuit and Evasion with Bounded
Control" IEEE Transacactions on Automatic Control, Vol. AC-25, No. 3, 1980.

[61] Shinar, J "Solution Techniques for Realistic Pursuit-Evasion Games", Advances


in Control and Dynamic Systems, C.T. Leondes, ed. Vol. 17, Academic Press, 1981.

View publication stats

You might also like