Download as pdf or txt
Download as pdf or txt
You are on page 1of 132

NATIONAL ORGANIZATION FOR MILITARY PRODUCTION

Science & Technology Center of Excellence

SHORT COURSE

IN

EXTERIOR BALLISTIC

By

Consultant Eng. ALHUSSIENY ALMOHANDES

CAIRO

September 2009

32
CHAPTER 1
GENERAL INTRODUCTION AND DIFINITIONS

Ballistics is the science, which studies the motion of


projectiles. Gunnery is the practical application of this science. By
nature ballistics is a mathematical subject; and in some of its
aspects it makes considerable calls upon the vast resources of
modern mathematics.

1.1 THE FOUR BRANCHES OF BALLISTICS:


1) Interior ballistic:
Interior ballistics deals with the operating processes within
the gun from the moment that the burning of the propellant is
initiated up to the moment it leaves the gun muzzle, or for rockets,
by the backward exhaust of the gas jet.
2) Intermediate ballistic:
Intermediate ballistics deals with the study of the transition
from internal to external ballistics, which occurs in the vicinity of
the gun muzzle and it is concerned mainly with the influence of
barrel oscillation. In the case of recoilless guns, the study is
extended to the region of the recoil jet nozzles.
3) Exterior ballistic:
Exterior ballistics deals with the description and
measurement of the motion of the projectile after it has left the
muzzle and following of its trajectory till it reaches its Target.
However, external ballistics also treats the trajectory of rocket-
assisted projectiles and rockets, and as a special case, the
dropping of a projectile in bomb ballistics.
4) Terminal ballistics:
Terminal ballistics may be defined as the study of the effect
of projectiles on a target. The condition under which missiles
impact against targets varies widely, depending on strike velocity,
strike angle, and the type of projectile and target.

33
During the period of exterior ballistics, the projectile is
subjected to the force of gravity, and the reaction of the air
resistance. Generally, the reaction of the air upon the projectile is
of complicated nature, comprising a number of forces and
couples. One force, however; called the drag; is so large in
comparison with the other forces and couples, that we may treat
its effect as of primary importance. The effect of the others can
then be considered as small corrections to this effect.
The drag is that force which directly retards the progress of
the projectile; it varies both in magnitude and direction during the
flight of the projectile. The consideration of the motion of the
projectile under this force and the force of gravity constitutes the
principal problem of exterior ballistics. The properties of this
motion are of the utmost importance upon a sound knowledge of
which success in gunnery largely depends.
To ensure the stability of an elongated projectile, it is
necessary to give it a spin around its axis; this spin tends to
counteract the tendency of the projectile to yaw under the
influence of the air’s reaction. The combined effect of the spin and
this reaction is that the nose of the projectile, in the case of a
right-hand spin, tends to point to the right of the vertical plane of
projection, thus given the rise to a lateral air reaction which
causes a bodily displacement of the projectile to the right, known
as drift.

34
1.2 DEFINITIONS:

Assuming a gun located at point (G), aiming at a target at


point (T), as shown in (Fig. 1.1), the gunnerists admit the following
definitions:

(Fig. 1): The parabolic trajectory of projectiles

1) Axis of the bore:


It is the line passing along the center of the bore. This may
be slightly curved. Generally the axis of the bore signifies the
direction of this axis at the breech.

2) Droop:
It is the vertical angle between the axis of the bore at the
breech, and the axis of the bore at the muzzle. This slight angle is
usually neglected.

35
3) Axis of the trunnions:
It is of the straight line passing through the center of the
trunnions, at right angles to the axis of the bore i.e. the transverse
axis about which the gun is rotated when elevating.

4) Caliber:
It is the diameter of the bore, excluding the depth of the
grooves.

5) Trajectory:
It is the curve described by the center of gravity of the
projectile in flight.

6) Muzzle (initial) velocity:


It is the velocity of the projectile at the muzzle of the gun.

7) Remaining velocity:
It is the velocity of the projectile at any point on the
trajectory.

8) Impact velocity:
It is the velocity of the projectile at the point of impact.

9) Line of sight:
It is the straight line passing through the sights and the
target (the line GT). It can also defined by the line joining the eye
and the target; when the eye is looking from the opened breech
through the bore to the target.

10) Vertical plane of sight:


It is the vertical plane that containing the line of sight.

36
11) Lateral plane of sight:
It is the plane passing through the line of sight, at right
angles to the vertical plane of sight.

12) Line of departure:


It is the direction of motion of the projectile at the moment
of leaving the gun. It is tangent to the trajectory at the muzzle.

13) Plane of departure:


It is the vertical plane containing the line of departure.

14) Jump (the angle J):


It is the vertical angle between the direction of the axis of
the bore at the muzzle before firing the round, and the actual initial
direction of departure of the projectile. When the line of departure
is above this reference direction, the jump is said to be positive
otherwise it is negative.
The following can be the mentioned as reference lines:
- The tangent to the axis of the bore in the region of the
muzzle, this gives the ballistic jump angle. This is
angle J that is normally used.
- The line joining the center of the muzzle and the center
of the chamber, this gives the apparent jump angle.
- The tangent to the axis of the bore in the region of the
chamber, this gives the virtual jump angle.

15) Angle of sight:


It is the angle, which the line of sight makes with the
horizontal plane. It is measured in the vertical plane of sight, and
is positive when the line of sight is directed above the horizontal
plane (the angle σ). When a negative angle of sight is given to a
gun it is denoted by depression.

37
16) Angle of departure (0f the projectile):
It is the angle, which the line of departure makes with the
horizontal plane (the angle α).

17) Angle of projection (of the projectile):


It is the vertical angle, which the line of departure makes
with the line of sight (the angle β).

18) Quadrant elevation (of the gun):


It is the angle, which the axis of the bore, when the gun is
laid, makes with the horizontal plane. It differs from the angle of
departure by the angle of jump (the angle Q).

19) Tangent elevation (of the gun):


It is the vertical angle, which the axis of the bore, when
the gun is laid, makes with the line of sight. It differs from the
angle of projection by the angle of jump (the angle Tang).

20) Point of impact (or arrival):


It is the point of intersection of the trajectory with the line
of sight (the point T).

21) Ballistic point of graze:


It is the point of intersection of the trajectory with the
horizontal plane through the gun (the point p).

22) Drift:
It is the lateral displacement of the projectile from the
plane of departure due to its rotation.

23) Summit (or vertex):


It is the highest point of the trajectory.

38
24) Angle of descent:
It is the vertical angle, which the tangent to the trajectory
at the impact point makes with the line of sight (the angle d).

25) Angle of arrival:


It is the vertical angle, which the tangent to the trajectory
at the point of impact makes with the horizontal plane through the
gun (the angle a).

26) Angle of incidence:


It is the angle, which the tangent to the trajectory at the
impact point makes with the normal to the surface struck (the
angle i).

27) Angle of impact:


It is the angle, which the tangent to the trajectory at the
impact point makes with the surface struck (the angle p).

28) CANT of the gun barrel (transverse tilt):


If the base of the aiming mechanism is not precisely
horizontal, or for a gun whose trunnion axis does not lie in the
horizontal plane, i.e. transversely tilted, an aiming errors arise in
azimuth and elevation, which can be considerable.
Due to the transverse tilt, the true vertical elevation of the
barrel is effectively reduced and the elevation aiming error thus
occurs.
In fact the error in azimuth is of greater significance than the
error in elevation in a gun, which is simply transversely tilted.
As a rule of thumb, at gun elevation of angle 45°, the azimuth
aiming error is approximately equal to the transverse tilt.

39
29) Time of the flight:
It is the time the shell takes to reach the point of impact
reckoned from the moment it leaves the muzzle.

30) Range:
It is the distance between the gun and the target.

Remarks:
1. The angle of departure α and projection β refers to the motion
of the projectile and not to the gun; therefore, they are called
the ballistic angles. While the quadrant elevation Q and
tangent elevation angle T refer only to the gun.
2. When the target is at the ballistic point of graze P, then the
point T will coincide with the point P, consequently the angle of
sight σ will vanish (σ = 0), and in this case, it will be:
- The angle of departure α will be equal to the angle of
projection β, i.e. (α = β).
- The quadrant elevation Q will be equal to the tangent
elevation T, i.e. (Q = T).
- The angle of descent d will be equal to the angle of arrival a,
i.e. (d = a).
3. Further if this target is in the horizontal plane we can deduce
that the angle of arrival a will be equal to the angle of impact p
and equal to the angle of descent d, i.e. (a = p = d).

The trajectory is a parabolic curve, on which any current point


as M as shown in (Fig. 1.2), will assume the elements (x, y, z),
the velocity v, the time of flight t, as well as the angle τ, through
which the horizontal at M directed towards the motion, must be
turned to coincide with the trajectory’s tangent, oriented in the
direction of motion:

40
Z

(Fig. 1.2): Projectile drift


The elements at the origin are suffixed by (o), at the summit by
(s), and at the impact point by (c), so at the origin we have:
x = xo = 0 , y = yo = 0 , z = zo = 0 , v = vo , τang = τo = α
Where: vo : is the muzzle velocity,
α : is the angle of departure.

The elements at the summit will be:


x = xs , y = ys , z = zs , v = vs , τang = τs = 0
Where: zs : in this case will measure the height of the trajectory.
Finally the element at the point of impact, on the assumption that
we are shooting against a horizontal target located at the point of
graze, will be:
x = xc , y = yc , z = zc , v = vc = vr , τang = τc = ω
Where: D : is the drift,
vr : Is the remaining velocity
x : is the range
ω : is the inclination at the impact point.

41
1.3 OBJECTIVE OF EXTERNAL BALLISTICS:
The problem that the combatants need always to solve is to
transport a projectile to a given point in the space, where it
develops its efficacy, in those conditions, which were fixed in
anticipation. Whatever this projectile might be small arms bullets,
artillery shells, aircraft bombs or auto-propelled projectiles.
The fundamental object of external ballistics is to enable the
combatant to reach his objective by the help of a collection of
numerical information’s condensed in a document, which we call
the Firing table of the concerned weapon.
On the other hand since the study of any projectile, starts by
the fixation of its efficacy, we can generally admit that external
ballistics can give light or the intrinsic characteristics of the
projectile (e.g. caliber, external profile, weight, etc) as well as the
initial and final conditions of firing.

The firing tables are calculated for well-defined conditions,


which we call the normal conditions being practically non
realizable and which refer to the so-called reference trajectories.

The passage from the normal conditions to the actual


conditions (preparation of firing) or the inverse passage
(interpretation of firing) is the permanent problems, which call
for the determination of the differential coefficients.

The principal object of the theoretical ballistics consists in


the reference trajectory calculation, as well as its concerned
differential coefficients. It depends mainly in this respect on the
physical data, which the experimental ballistics offers.

Meanwhile the applied ballistics is concerned with the study


of the apparatus and methods of firing in the shooting ranges, as
well as the construction of the firing tables.

42
1.4 THE ANGULAR MOTION OF PROJECTILES:
As a sequence of the rifling of the gun barrel, the projectile on
leaving the muzzle will possess a very high velocity of rotation,
which we call the "spin". The' clearance existing between the
projectile and the bore, submits the projectile at the muzzle to a
sort of transversal percussions. Such percussions combined with
the spin results in what we call the motion of "nutation", during
which the nose of the projectile will be describing a sort of
sinusoidal vibrations.
The nutation is fortunately amortised after a flight of few
hundreds of meters, such circumstance permits the precision of
firing.
The slight angle enclosed between the projectile's axis and
the muzzle velocity direction, at the instant of leaving the muzzle,
is called the "initial angle of yaw". It varies along the trajectory.
The projectile flying in the air with its C.G. tracing its
trajectory, meanwhile its axis is inclined to the direction of the
tangent of the trajectory by the "angle of yaw”, will be subjected
to the air resistance, which is assumed to be acting at a certain
point on the projectile axis; namely the "center of resistance”.
The resultant action of the spin, the air resistance, as well as
the curvature of the trajectory is a motion of "precession".
Precession is the state when the nose of the projectile is
describing a circle round the tangent to the trajectory, meanwhile
remembering that its C.G. is always lying on the trajectory, or in
other wards the line "C.G. - Nose" will be describing a cone
round the tangent of the trajectory, having an angle being twice
as much as the current "angle of yaw".
The combination of the precession and the still alive nutation
in the neighbourhood of the muzzle, is schematized by a
continuous vibratory sinusoidal movement rolling and embedded
in the circular periphery of the previously, mentioned cone.

43
Finally, for a gun rifled to the right, the precession motion is
not exactly symmetrical about the tangent of the trajectory, but it is
slightly deviated to the right from the plane of firing.
This continuous deviation leads to the final "drift" at the
impact point. The drift is generally small as compared with the
projectile’s range.
As a result of the above mentioned description of motion
considering both nutation and precession motions as simple
"perturbations"; together with the fact that the drift is small; we
can treat the "principal ballistic problem" as the motion of the
C.G. under a tangential air resistance, or in other words, we
assume that the trajectory is completely contained in the plane of
firing.
The projectile is subjected to a system of forces, some of
which are known, while the other unknowns may be either
approximately measurable, or even non-obedient to any kind of
measurement. Among those unknowns we can point out the
previously mentioned angular motion.
Consequently we need to be contented with the measurable facts
on the trajectory. Such reliable measurements may be used for
the analysis of this system of forces in an inductive manner, and
in such a way one may think of creating an almost completely
empirical ballistics. Such trials have been accomplished by some
ballisticians and ended in complete failure, simply because the
facts appear isolated and no reciprocal checks or controls can be
achieved

44
1.5 PHYSICAL ASPECT OF AIR RESISTANCE:
When a. projectile flies in the air, the air in the neighbourhood
of the path of the projectile is set into motion. As the projectile
pursues its course more and more air is set into motion. The
energy thus gained by the air can only be drawn from the energy
of the projectile. There is therefore a continual drain upon the
projectile's energy, with the consequent result that it continually
loses velocity in its flight.
At the nose of the projectile, the air undergoes compression.
The air being an elastic fluid, compression formed at any point in it
will be transmitted in all directions with a velocity, which is in
general, the velocity of sound.
If now the projectile is travelling with a velocity less than that
of sound, the compression of the air at the nose will be
transmitted; as soon as it is formed; away from the nose in all
directions.
If on the other hand, the projectile is travelling faster than the
sound velocity, the compression of the air at the nose cannot be
transmitted away from the nose in all directions, it can only be
transmitted away laterally but not forwards. The result is that the
projectile's nose will always be in contact with a cushion of
compressed air.

Finally, at the base of the projectile, the air undergoes


rarefaction. The pressure of the air at the base is therefore
reduced, and the effect of the pressure on the nose is therefore
increased. The surrounding air, in the endeavour to equalize the
pressure behind the projectile, moves rapidly inwards towards the
base and so forms a train of "eddies" behind the projectile.
When we photograph a projectile moving faster than sound,
we can distinguish the nose and tail "shock waves" as well as the
zone of eddies.

45
The air resistance is the only unknown force that figures in the
principal ballistic problem, for which, an exact analytical
expression have not yet been determined. This fact explained why
the artillerists had always rushed towards the practical
experiments for the numerical determination of the value of air
resistance.
The simplicity of the experimental determination of the air
resistance by direct firing enabled the ballisticians, since the early
times, to compile the relation between the air resistance and the
projectile's velocity in the form of tables or graphs or even to
derive an empirical expression which represents this relation for
any given range of velocity, evidently for those types of projectiles
on which they experimented.

46
CHAPTER 2

MOTION IN VACUUM

As mentioned before, external ballistics deals with the


description and measurement of the motion of a body, which is
projected at an angle θ°, and with an initial velocity vo.

In a narrower sense, external ballistics is understood to be the


following of the trajectory, which unpowered projectile, fired in
the earth's gravitational field where the attraction of the earth
affects upon the projectile, as well as the air resistance. If the
motion takes place under just the influence of the acceleration due
to gravity, then it speaks of vacuum ballistics.

However, external ballistics also treats the trajectories of


rocket-assisted projectiles and rockets and, as a special case,
the dropping of a projectile in bomb ballistics.

Thus, if the whole process is divided up, external ballistics


comes between intermediate ballistics and terminal ballistics.

Once the projectile has left the gun and the influence of
emerging gases has finished; the part of the flight that known
as external ballistics begins.

There are a number of factors, which affect the motion of the


shell, some associated with the shell itself and others
associated with the medium (air) through which the shell is
moving.

The properties of the shell that enter into the problem are
mass, caliber, nose shape, spin rate and velocity, whilst those
47
associated with the medium (atmosphere) are medium (air)
density, temperature, pressure and viscosity.

One of the prime considerations in shell design is accuracy


of fire and so some means of stabilization must be employed.
The two most common means of achieving this are spin and
fin stabilization.

If the shell is spun the combined action of gyroscopic and


aerodynamic effects will cause it to drift sideways. Another
significant effect for longer ranges is that due to the rotation
of the earth. We consider these factors in more detail later. We
begin our study of external ballistics by considering the
motion of a projectile under the influence of gravity only, that
is, we assume that the projectile is fired in a vacuum.

48
2.1 MOTION UNDER NO EXTERNAL INFLUENCES:
Consider a projectile fired at an angle of elevation è with a
muzzle velocity V meters per second (m/s). Newton's first law of
motion states that a body in motion continues to move in a straight
line at a constant velocity V unless acted upon by an external
force. Therefore, in the absence of gravity or other forces, the
projectile would continue in its initial direction and maintain its
muzzle velocity as shown in the following (Fig. 2.1):

(Fig. 2.1): Motion of a projectile under no external influences

But, of course, there is an external force, the force due to the


Earth's gravitational field, which has the effect of pulling the
projectile back towards the center of the earth with an acceleration
of g (m/s2). The value of g varies with distance from the earth, but
for short range weapons, such as small arms and field guns, it can
be assumed that the gravitational field is uniform, and take a
constant value 9.81 (m/s2).

49
2.2 MOTION IN VACUUM UNDER INFLUENCES OF GRAVITY:
As a start, If in the simplest form, the gravitational
acceleration g = 9.81(m/s2), {more precisely, its exact value will
be: g = 9.780490 (1 + 0.0052884 sin2 β – 0.0000059 sin2 2β)
where “ β “ is the geographical latitude}, is only considered, and
the curvature and the rotation of the earth are ignored, then
results can still be obtained, which in some cases, come close to
reality. These calculations are particularly useful for rough
approximations. By neglecting air resistance, however, they lead
to overestimation of range.
2.2.1 The Trajectory:
The trajectory can be computed from Newton's second
law of motion in the x and y direction, as follows:
m (d2x/dt2) = m x  = o (1)
m (d2y/dt2) = m y  = -mg (2)
Integrating these equations with respect to time, and using
the initial conditions that at t = 0, one obtains the coordinates of
any point on the trajectory at a time t:
x = 0, vox = vo cos θo,
y = 0, voy = vo sin θo
vx = x = dx/dt = vo cos θo
x = vo (cos θo) t (3)
t = x / vo cos θo (4)
vy = y = dy/dt = vo sin θo - gt
y = vo (sin θo) t – ½ gt2 (5)
where: t : is the flight time to a point on the trajectory (x, y)
θo : is the initial angle of departure of the projectile
vo : is the initial velocity of the projectile
m : is the mass of the projectile

50
After eliminating the time t, the equation of y becomes:
y = x tan θo – ½ g (x / vo cos θo)2 (6)
which is a parabola that is symmetrical about the ordinates yG of
the apex, as shown in (Fig. 2.2):

(Fig. 2.2): Parabolic trajectory of projectile in vacuum


from equation (6), the
total horizontal range will be at y = o:
X = 2 (vo2/g) sin θo cos θo
= (vo2/g) sin 2 θo (7)
i.e. the maximum
horizontal range is obtained at: θo = 45°, so:
Xmax = vo2/g (8)
and the total time of
flight will be obtained from equation (4):
T = 2 (vo/g) sin θo (9)
and the vertex height
will be:
yG = (g / 8) T2 (10)

51
This expression yG, known in the literature as Haupt's
equation, is only exact in a vacuum, but can also be used in
the atmosphere as a good approximation for estimating the
vertex height yG.

2.2.2 Parabola of Safety:

If, for a given initial velocity vo, the angle of departure θo is


varied, one then obtains a family of parabolic trajectories with the
enveloping curve symmetrical about the Y-axis, as shown in the
following (Fig. 2.3):

y = (vo2/ 2g) – (gx2/ 2 vo2) (11)

(Fig. 2.3): Parabola of safety

Outside of this so-called parabola of safety, no target can be


hit with the vo specified. Each point on the enveloping curve can
be reached by only one trajectory. However, all points inside the
enveloping parabola can be reached by two trajectories. Targets
lying below the trajectory for angle θo= 45° can be reached both
by low angle fire (θo < 45°) and high angle fire (θo > 45°). For
targets on the horizontal line the sum of departure angles θo of
these two trajectories is 90°.

52
2.2.3 Firing on an Inclined Plane; Lifting the Trajectory:

If the line to the target makes an angle γ with the horizontal


line, we have for the range:

Xγ = (2vo2/g) [cos θo sin(θo- γ)/ cos2 γ] (12)

and for the flight time:

T = (2vo/g) [sin(θo- γ)/ cos γ] (13)

In order to hit the target A, as shown in (Fig. 2.4), the


trajectory must be lifted, i.e., the super-elevation angle θ1 must be
added to the angle of site γ, i.e. the overall angle of elevation will
be:

θo = γ + θ1 .

(Fig. 2.4): Firing on an inclined plane

For sloping ground (uphill or downhill), the same super-


elevation angle is used as for level ground, then a round will
normally fall short for positive site angles, and overshoot for
negative ones. However, for small site angles γ, this error is small.

53
2.2.4 Trajectory in Vacuum by the Fishing Rod Principle:

Using the "fishing rod principle" as shown in (Fig. 2.5), the


trajectory can be determined as the resultant of a uniform motion
in direction of the initial departure elevation, and a uniform vertical
acceleration. If Q - P is established for a set flight time t, so that x*
and x make the angle θo, then for a muzzle velocity vo and
elevation θo, P gives the position of the projectile after t seconds.

(Fig. 2.5): Trajectory determination by the "fishing principle"

Lifting the trajectory, in accordance with the "fishing rod


principle", gives valid approximations also under atmospheric
conditions for small angles of elevation and lifting. It is a better
method than that given previously for the lifting of the trajectory.
By changing the angle θo (Fig. 2.5), which corresponds to the
elevation θ1 (Fig. 2.4), i.e. by lifting the "fishing rod" up and
down, any trajectory can be determined.

54
2.2.5 The Beaten Zone:

The beaten zone, as shown in (Fig. 2.6) is understood to be


that distance Δx in the target area, in which the flight altitude y is
less than or equal to the target height Δy, The following
considerations are interesting because, for a given target height
Δy, they give a clue as to how precise the range measurement x
must be, and how great will be the hit probability which is
dependent on Δx.

(Fig. 2.6): The calculation of the beaten zone

The general trajectory can be represented quite well by


parabolic curves. This is used to advantage for the last section of
the curve. From the vacuum trajectory Equation (9), one obtains
the following relationship:
2
Δ x = (vE /g).sin ω.cos ω - K (14)

Where:
K = {(vE /g) cos ω [(vE /g) sin ω - 2Δy]}1/2
2 2 2 2

ω = Impact ( Incident) angle of the projectile


vE = The velocity at the target

55
By using the vertex altitude YG {Equation (10)}, this can be
considerably simplified to:
1/2
Δ x = 2 YG cot ω {1 – [1- (Δy/ YG)]} (15)

For the case where Δy<< YG, which frequently occurs in


practice, this becomes:

Δ x = Δy cot ω (1 + Δy/ 4 YG) (16)

or, with the range x:

Δ x = Δy cot ω (1 + Δy cot ω / x) (17)

These equations can be applied quite well under atmospheric


conditions, if, instead of 4YG in Equation (16), one uses the
quantity 3YG (DUFRENOIS); Equations (16) and (17) then
become the Equations:

Δ x = Δy cot ω (1 + Δy/ 3 YG) (18)

Δ x = Δy cot ω (1 + 4Δy cot ω / 3x) (19)

or, based on the investigations of W. KRAUS:

Δ x = Δy cot ω (1 + 3Δy cot ω / 2x) (20)

56
2.2.6 The Characteristics of "In Vacuum" Trajectories:

We now have the situation shown in (Fig. 2.7) for the


motion of a projectile under the action of gravity:

(Fig. 2.7): Motion of a projectile under the action of gravity

1. The trajectory (The trajectory is the path taken by the


center of gravity of the projectile) is symmetrical about
the vertical line through the apex of the trajectory and
moves along a curve called a parabola.
2. The whole trajectory lies in the vertical plane containing the
line of departure.
3. The range depends upon the muzzle velocity and increases
as the angle of departure θ° increases up to 45°. For a
given velocity, the maximum range is achieved at 45°.
4. As θ° increases above 45°, the range decreases until it
becomes zero at θ° = 90°. At this stage, the maximum
height is reached and also the time of flight reaches its
maximum value.

57
5. The angle of arrival equals the angle of departure and the
final velocity at the point of impact equals the muzzle
velocity. The velocity is a minimum at the vertex.
6. The trajectory is entirely independent of the nature of the
projectile, since air resistance is zero.

Although the in-vacuum theory is relatively simple, the


mathematics involved in the study of the real trajectory is much
more complicated. However, it is worth noting that it is accurate
for ballistic trajectories that are predominantly outside the
Earth's effective atmosphere i.e. for very high altitudes, such
as those of Inter-Continental Ballistic Missiles {ICBM's), and is
reasonably accurate for small projectile velocities, or large
projectile weights such as mortars if the muzzle velocity is
less than 250 m/s.

58
CHAPTER 3

METEOROLOGICAL CONDITIONS

3.1 PROPERTIES OF THE ATMOSPHERE:


The atmosphere up to 20 km altitude is basically composed
of 78% nitrogen, 21% oxygen with the remaining 1 %
comprising water, carbon dioxide, hydrogen and, other
gases. At very high altitudes, separations of the gases occur. For
example at about 100 km, atomic oxygen separates with the
lighter gases diffusing upwards and separating out into further
layers.
Pressure, temperature, density and viscosity all vary with
altitude. These continual changes in the physical properties of the
atmosphere affect the resistance of the air and therefore the
range of a projectile. Since the trajectory of small arms, artillery
and ballistic missiles have peak altitudes typically of 50 m, 20
km, and 600 km respectively; projectiles fulfilling these roles will
experience a significantly different environment during their flight.
For example, the density of air decreases as the height above
sea-level increases and consequently the range of a missile
increases as the trajectory gets higher and higher. If the
projectile is fired into the upper atmosphere, where the air
resistance is very small, it will go much further. The Germans
who launched the “V2” rockets during the Second World War used
this principle. Once the “V2” was about fifty miles up, there was
practically no resistance at all. On the other hand it is the de-
crease of density with altitude, which causes some limitations on
the performance of certain aircraft; for example, the helicopter
has a maximum operating ceiling due to lack of lift.
The effects of variations in atmospheric conditions depend on
the point of the trajectory at which they operate. To predict
trajectories the air through which the projectile passes is divided
into layers and a weighting factor is determined for each layer.

59
Weighting factors vary not only from layer to layer but also
from one trajectory to another. To achieve the greatest accuracy
there should be a weighting factor for each layer of each
trajectory: in reality this is only practical when calculations are
performed on a computer.
The definition of a standard atmosphere is essential to the
ballistician as this enables him to undertake stability calculations
and range estimations under normalized conditions, rather than
for a general situation. It is also desirable to have this standard
atmosphere close to some average value of conditions. The most
commonly used standard atmosphere is that used by the
International Civil Aviation Organization (ICAO). Other
standards include the World Meteorological Organization
Standard, the International Organization for Standardization, ISO
Standard. Sort of typical figures given in such atmospheric tables
are shown in (Table 3.1):

(Table 3.1): Variation of P, , T and  with altitude


Po, ρo, To and o refer respectively to the values of the pressure,
density, temperature and viscosity at sea level.

60
3.2 THE BASIC PHYSICAL PROPERTIES OF GASES:

 Density ( ρ ):

The density of a gas is its mass per unit volume. The density
of the standard atmosphere at sea level is:
ٌρo = 1. 23 (kg/m3)

 Temperature (T):
The temperature is a measure of the average kinetic energy
of the gas molecules. The Kelvin scale of temperature
adopts zero K as absolute zero with Celsius degree
intervals. The American absolute scale of temperature is
Rankin (°R), and this scale has Fahrenheit intervals. They
are related as follows:
°
R = °F + 460
K = 5/9 °R
It will be noticed from (Table 3.1) that the temperature does
not decrease uniformly with altitude.

 Pressure (P):
The pressure of a gas can be related to the density of the gas
through the perfect gas equation, which may be written in the
form:

P=ρRT

Where: R is a constant for a given gas and in the case of air, it


has the value of 0.287 (kJ/ kg. K).
Note: that when the temperature remains the same, the
pressure increases directly with the density and vice-
versa.

61
 Absolute Viscosity:
As the pressure on a gas increases the gas becomes denser.
Generally speaking, when a fluid, that is, a liquid or a gas,
becomes denser, then it becomes “stickier”, or more viscous.
The important difference between the viscosity in a liquid and
a gas is that for a liquid, the viscosity can only increase if it is
cooled down, while the viscosity of a gas can vary with the
temperature, or the pressure, or both. In fact, the viscosity
of a gas rises with an increase in temperature.

 Kinematical Viscosity:
It is a further important property of fluids used in ballistics and
aeronautics, which is defined to be the absolute viscosity divi-
ded by the density i.e.:
Kinematical Viscosity = Absolute Viscosity / Density
The term viscosity is usually used to mean kinematical
viscosity. The increase in viscosity on a hot day is offset
by the reduction in overall drag owing to the reduction in
air density, so the range is generally increased.

3.3 WIND EFFECT:


The effect of wind on a particular projectile depends very
much on the type of projectile fired. In general a head wind will
impede the passage of a shell and a following wind will assist it.
A cross-wind blowing from left to right will cause the shell to
deviate to the right and vice-versa.
A spinning shell is not affected as much as a finned projectile
e.g. a rocket, because its lateral wind resistance is smaller. Wind
effects are very important for rocket trajectories, especially
for high angles of fire. If the wind is blowing across the trajectory
it blows on the fins of the rocket, and as these are at the base, the
rocket nose will turn into the wind.

62
3.4 THE SPEED OF SOUND IN AIR (a):
The speed of sound in air can be expressed directly in terms
of the temperature of the air as follows:

a = ( γ R T )1/2
Where:
γ : is the ratio of specific heats,
For air at atmospheric pressure: γ = 1.404
R: is the gas constant,
For air: R = 0.287 (kJ/ kg. K).
If the velocity of the projectile exceeds that of sound the
projectile becomes supersonic. From the above relationship, it is
evident that the velocity at which transition occurs is directly
related to the temperature of the air.

3.5 ROTATION OF THE EARTH


When dealing with long ranges the effects due to rotation of
the earth must be taken into account. The rotational speed of the
earth's surface at the equator is about 450 m/s, i.e. 1620 km/h. At
first sight this appears to be of no consequence as the gun, target
and atmosphere are all travelling together, more or less, at the
same speed. However, the motion of the earth's surface is, not
in a straight line, so during the period that a projectile is in
flight the target may be carried sideways, up or down by the
earth to which it is attached. Fortunately the rotation rate and
dimensions of the earth are fixed, so for any gun and target at
known positions on the earth's surface it is possible to apply
corrections to the gun bearing and elevation to ensure that
the projectile meets the target. A rigorous treatment of this drift
is beyond the scope of this study. However, it is not difficult to get
an adequate grasp of the principals involved in the following
chapters.

63
CHAPTER 4

AIR RESISTANCE

4.1 THE COMPONENTS OF AIR RESISTANCE:

The Components of Air Resistance that affecting the


projectiles during its motion in the medium of the atmosphere are:

1. Fore body Drag:


a. Compression Waves
b. Shock Waves
2. Base Drag
3. Skin Friction
4. Excrescence Drag

64
4.1.1 Fore body Drag:
a. Compression Waves:
As the projectile moves through the air, the air is
displaced. The energy needed to displace this air causes a
continual drain on the initial kinetic energy of the projectile: this
loss of energy is called the drag and causes a continuous loss of
projectile velocity. The compression of the air, which occurs
immediately in front of the projectile, is transmitted to the
surrounding air as a pressure wave. In its turn this causes a
disturbance that travels through the air at the same speed as
sound waves, since sound waves are themselves merely
disturbances of the air. When the projectile is travelling at a speed
below that of sound, that is below approximately 340 m/s, the
disturbances move faster than the projectile and so spread out
away from it, as shown in (Fig. 4.1):

(Fig. 4.1): Compression waves at subsonic speed

The generation of these compressed waves causes


the projectile to experience "fore body drag". Fore body drag,
also, is of greatest significance in the supersonic region.

65
b. Shock Waves:
When the projectile is travelling faster than sound no
part of the disturbance can escape directly in front of the
projectile since it is moving faster than the disturbance. The
result is that the compression waves coalesce or "bunch up",
and a shock wave is created at the nose of the projectile, as
shown in (Fig. 4.2):

(Fig. 4.2): Shock wave at supersonic speed


In general, a conical shock wave is generated which has
an angle θ°:

sin θ° = 1 / M

Where: M is the Mach number that is defined as the


velocity of the projectile V divided by the local speed of sound in
air, a, that is:
M =V/a

sin θ° = a / V

Which show how rapidly the shock cone decreases with


increase of Mach number.

66
At hypersonic speeds, which defined as over Mach 5,
i.e. (M > 5), the shock wave almost follows the shape of the body.
With sharp-edged or pointed bodies, the shock wave usually
originates from the nose of the body, so the shock wave is
attached, as shown in the following (Fig. 4.3):

(Fig. 4.3): Shadowgraph of 7. 62 mm bullet

If the body has a blunt nose or a larger angle, then the


shock wave forms just ahead of the body, so the shock wave is
detached, as shown in (Fig. 3.4):

(Fig. 4.4): Shadowgraph of high drag training round

67
The more blunt the body, the higher the shock wave drag
and therefore the greater the retardation. This is clearly shown in
the flight of space vehicles where the exit profile shape has an
ogival or conical cover, which is jettisoned before reentry to reveal
a blunt fore body which can dissipate the consequent high thermal
energy and give the necessary retardation, as shown in (Fig. 4.5):

(Fig. 4.5): Shadowgraph of space re-entry model


(Detached shock wave)

Because we are dealing with finite length bodies


travelling supersonically, the disturbance generated must also be
finite, and so a bow or nose shock wave is accompanied by a
stem or tail shock. This pair of shock waves constitutes the
"sonic boom" heard from bullets and shells as well as from
supersonic aero planes such as 'Concorde'.

68
4.1.2 Base Drag:

It is clear from the shadowgraphs shown above that there is


considerable turbulence behind the bodies. This turbulence called
the wake causes a further resistance known as "base-drag". The
cause is a region of low pressure immediately behind the
projectile that results because the air flow cannot return quickly
enough to fill the space behind the projectile. The consequence is
a vacuum or suction effect that shows itself in the form of a
resistance to motion, as shown in (Fig. 4.6):

(Fig. 4.6): Base drag

69
4.1.3 Skin Friction:
Additional resistance to motion is caused by air adhering to
the surface of the projectile: it is known as "skin friction". The
mechanism is basically that the air at the surface of the shell is
moving at the same speed as the shell; the next layer of air is
moving a little more slowly, and so on outwards. Frictional drag,
which results, is generally relatively small for most shells; but for
larger missiles, the effect becomes more important and certainly
has to be considered in rocket calculations. A smooth and
polished surface will help to reduce skin friction. The volume in
which these viscous forces act is called "the boundary layer".

4.1.4 Excrescence Drag:


The final form of resistance called "excrescence drag"
arises from protuberances from the shell and can be minimised
by eliminating all unnecessary projections on the projectile.
Normally, only the driving band is unavoidable and this must be
carefully designed.

70
4.2 THE SHARE OF MAIN CAUSES OF AIR RESISTANCE:

The following diagram (Fig. 4.7) is for a conventional


projectile and shows the relationship between the three main
causes of air resistance and the projectile velocity:

(Fig. 4.7): Air resistance

71
4.3 VARIATION IN DRAG VALUES:
Fore body drag increases steadily as velocity increases, and a steep rise is noticed as the
velocity of sound is approached. The increase is maintained, for a time, at this higher rate in
the supersonic zone, but gradually reduces. Base drag increases with velocity until the velocity
of sound is reached, but then remains fairly constant. This is because, as the velocity of the
projectile approaches the speed of sound, the air pressure behind the base tends to zero, but
since the pressure cannot become absolutely zero, any increase in velocity has little effect.
Skin friction is normally of least consequence but as mentioned earlier, it can be relatively
greater for long thin projectiles.

At this stage it is worth commenting on the transonic zone.


Around the velocity of sound there is a zone in which the
projectile's behaviour is unpredictable because the air resistance
is changing extremely rapidly. This is noticed in the previous
diagram, (Fig. 4.7), especially as regards fore body drag. In this
transonic zone small changes in the projectile's velocity, due
perhaps to wind fluctuations, cause marked changes in
resistance. It is important that the projectile is steady when it
enters the transonic zone. Since it is common for projectiles to be
unsteady just outside the muzzle of the gun, it is undesirable to
use a charge, which gives a muzzle velocity only slightly above
the velocity of sound. For this reason, the designer if possible,
avoids muzzle velocities in the region of the velocity of sound.

72
4.4 THE FORM OF THE REAL TRAJECTORY IN AIR:
The main effect of air resistance is to impede the progress of
the projectile, both in the rising and falling parts of the trajectory,
and to distort the in-vacuum trajectory shown in (Fig. 2.7). Air
resistance can be considered as a retarding force always acting in
a direction opposite to the motion of the projectile's velocity. For
the in-vacuum trajectory, the vertical component of the projectile's
velocity is retarded by gravity, while the horizontal component is
constant.
The in-vacuum trajectory and the real trajectory in air for the
same projectile and the same initial flying conditions are as shown
in (Fig. 4.8). In air, during the rising portion of the trajectory, air
resistance acts in addition to gravity, and the vertical component
of the velocity is reduced to zero more rapidly. Air resistance also
reduces the horizontal component of the initial velocity:

(Fig. 4.8): Real and in-vacuum trajectories

111
During the fall the air resistance acts to oppose gravity so that the time for the fall is
greater than the rise. All these effects can be summarized as:

1. The trajectory is no longer symmetrical: the vertex is nearer to


the point of impact than to the gun.
2. The vertex height is less than for the corresponding "in-vacuum” trajectory.

3. The angle of descent is greater than the angle of departure.


4. The terminal velocity is less than the muzzle velocity.
5. The smaller the drag coefficient of the projectile the nearer
point A is to B.

Before studying the drag coefficient we consider the results in


the following table, (Table. 4.1) which show that the presence of
the atmosphere modifies considerably the maximum range
predicted by the in-vacuum theory:

Type of Round Muzzle Velocity Maximum Range


(m/s) In- Vacuum In- Air
300 mm Mortar 396 16 km 11 km
155 mm FH70 700 49 km 24 km
7.62 mm SLR 840 70 km 4 km

(Table. 4.1): The effect of air resistance on certain projectiles

112
4.5 THE DRAG COEFFICIENT:
The drag coefficient is a convenient way of expressing the
ballistic performance of a particular projectile. We have seen that
aerodynamic drag can be broken down into a number of
components; in general the total drag force can be expressed as:

Drag force = (1/2) ٌ V2 A CD

Where ٌ is the density of air, V is the velocity of the


projectile, A is the cross-sectional area based on the caliber of
the projectile, and CD is the drag coefficient. CD has a value
between 0 and 2 that is dependent upon the streamlining of the
projectile and its attitude in flight.
Typically a shell shape gives a value of CD = 0.3 while for a
blunt body like a training round CD = 0.8.
In general the lower the value of CD the lower the drop in
velocity with range. It must be remembered however that CD is
only constant at speeds well below 300 (m/s) or well above 1500
(m/s). When CD is constant, it means of course that the drag is
proportional to V2. The variation of CD between 300 and 1500
(m/s) is discussed later.

113
4.6 THE BALLISTIC COEFFICIENT:
In recent years it was commonplace to speak of a "ballistic
coefficient" when describing the ability of a projectile to fly a
particular trajectory or to be compared with the trajectory of
another projectile. The term standard ballistic coefficient was
defined as:

CO= W/ Kٌ d2

Where W was the projectile weight in (lbs), Kٌ was an


adjusting term for "shape and steadiness" with a value close to
unity, and d is the caliber of the round in (inches).

In practice CO took a value of between 2 and 3 for a


reasonably efficient shell, and it could be seen that if the ratio of
W/d2 was increased then the shell went further for the same
muzzle velocity, The value of CO was used in conjunction with the
special tables called P(v) tables to predict the retardation of a
particular shell shape at different velocities.

However, we now use CD, in line with modern aerodynamic


terminology.

114
CHAPTER 5

REDUCING THE AIR RESISTANCE

5.1 METHODS OF REDUCING THE AIR RESISTANCE:


Methods of reducing the air resistance on the projectile during
its motion in the medium of the atmosphere are:
1. Boat tailing
2. Nose shape
3. Base bleed

5.1.1 Boat Tailing:


Aerodynamic drag has a significant effect on the range and
performance of high-speed projectiles, so it is desirable to design
a projectile with the minimum drag. It may cause other problems
such as instability but, as in most design problems, a "trade off"' is
required. Below the speed of sound base drag and skin
friction are the dominant causes of resistance. To reduce base
drag, boat tailing is often used. A boat tail is usually a truncated
cone of the order 1/2 - 3/4 caliber in length: it reduces the
pressure drop in the wake of the projectile by allowing the air to
flow in and around at the base, as shown in (Fig. 5.1):

(Fig. 5.1): Boat tail design

115
The optimum value of the angle θ° has been found to be
about 7.5°. Recently new shapes have been proved to be superior
to conical both from the base drag aspect as well as aerodynamic
stability at transonic speeds. These shapes are formed by slicing
planes at an angle to the cylindrical body in the base regions.
They result in square and triangular base shapes.

Tests have shown that a longer range is acquired with boat


tailing than a flat base projectile of the same weight and velocity.
Boat tailing is undoubtedly effective at subsonic speeds and in
fact it was originally designed to improve the long-range subsonic
performance of machine guns.

However, there is a difference of opinion concerning its


effectiveness at supersonic speeds; reports that an increased
wake dispersion results when projectiles are supersonic, with a
subsequent decrease of performance. We know that there are
disadvantages of boat tailing due to manufacturing problems,
greater barrel wear and erosion. On the other hand provides us
with results that show that there is a substantial increase in range
because of the boat tail effect. It is found on many supersonic
projectiles, but a complete justification of its use has not yet been
formed.

116
5.1.2 Nose Shape:
The air resistance suffered by a shell in flight depends on
the shape of the shell; of particular importance are the shapes of
the nose and the base.

a) The Old Caliber Radius Head shell:


Prior to World War I, the shape of the nose was usually a
simple ogive as shown in (Fig. 5.2): where: AC is an arc of a
circle drawn about O as center:

(Fig. 5.2) Caliber radius head

. If OA = n x AB, the shell is said to be of n (c r h); calibre


radius head. Normally n would be about 2.

117
b) The Modern Shell Head:
They are normally given a more pointed nose of
somewhat different shape. The point C is located as before but
the actual profile between A and C is an arc of longer radius
drawn about a new point P as shown in (Fig. 5.3):

(Fig. 5.3) Fractional caliber radius head

AC is bisected at R, and the line RO is produced in the


direction of P. A center of curvature S can be selected anywhere
on OP. With S as center, an arc AC of radius SA is struck, and it
is this, which defines the semi-profile of the head. A head of this
kind is defined as a fraction. Thus a 5/10 (c r h) caliber radius
head is a head the length of which corresponds to that of a simple
ogive of radius 5 caliber. The curvature is 10 caliber i.e. SA is
ten projectile diameters.

118
Sub-sonically, fore-body drag can be significantly reduced by
any suitable smooth nose shape. The graph at (Fig. 5.4) shows
the relationship between the drag coefficient CD with an increase
in speed mainly in the subsonic region, for three various shapes of
nose:

(Fig. 5.4): The variation of CD with speed in the subsonic


region for different nose shape projectiles

At supersonic speeds, the effect of base drag and skin friction components are similar to
those experienced below the sound barrier; but the shape of the projectile has a great influence
on the fore-body drag. Mathematical methods are available to calculate the shape of least
resistance of a projectile if the dimensions are known.

119
In general, the longer the head the smaller the fore-body drag
at supersonic velocities. The total length of a spin stabilized
projectile lies between 4.5 and 6 calibers of which between 2.7
and 4 calibers are available for the head. It must be pointed out
however that for important technical reasons, the total length has
generally to be restricted to 5 calibers. Only if tactical reasons
absolutely required an exceptionally thin projectile would this
value be exceeded.

As shown in (Fig. 5.5), the variation of the drag coefficient CD


with speed for high velocities, for two differently shaped projectiles
is illustrated:

(Fig. 5.5): The variation of CD with speed for two differently shaped projectiles

120
Also (Fig. 5.6) shows the variation of CD with speed for a
typical shell:

(Fig. 5.6): The variation of CD with speed for a typical shell

121
5.1.3 Base Bleed:
A practical method of reducing drag, especially for larger projectiles is "base bleed".
This process involves burning a small quantity of propellant, which is fitted into the rear of the
projectile. The propellant burns at low pressure and generates a jet of gas at the base, which in
turn increases the pressure behind the projectile base and can reduce base drag by up to 50%.
Range increases of between 10% and 30% have been achieved with this method. The graph
below; (Fig. 5.7) illustrates the marked effect of base bleed On a 120 (mm) experimental
projectile:

(Fig. 5.7): The effect of base bleed on range

5.1.4 External Burning:

It is a more sophisticated version of base bleed, which emits


more gas through holes in the periphery of the base.

122
CHAPTER 6

STABILITY OF PROJECTILES

6.1 YAW OF THE PROJECTILE:


Except by chance, when a projectile is fired from a gun its axis
does not lie exactly along the trajectory, which is the path of its
center of mass.

The angle between the axis of the projectile and the


tangent to the trajectory is known as the angle of yaw, as
shown in (Fig. 6.1):

(Fig. 18): Angle of yaw

(Fig. 6.1): The angle of yaw

6.2 FORCES ACTING UPON THE PROJECTILE DURING ITS


FLIGHT THROUGH THE MEDIUM OF ATMOSPHERE:
1) Drag
2) Crosswind Force
3) Over-Turning Moment:

The two major aerodynamic forces acting upon the


projectile, which affect and depend upon yaw are drag and
crosswind force.

123
6.2.1 Drag Force:

Drag retards the motion of the projectile along its trajectory


and varies in magnitude with angle of yaw. During normal flight
a projectile inevitably yaws slightly, and this produces a relatively
small significant increase in drag.
If the projectile yaws to greater angles by design, failure, deflection or encounter with
a dense medium (e.g. water), the drag will increase dramatically. For example: if a supersonic
projectile in air yaws to 90° its drag typically increases a hundredfold.

6.2.2 Cross-Wind Force:

As shown in (Fig. 6.2), Crosswind force is similar to the lift


force generated by a wing. This force acts perpendicular to the
trajectory and in the direction of yaw. It is of zero magnitude at
zero yaw, and increases with increasing yaw within the limited
angles of yaw occurring during normal flight. At extreme angles of
yaw the crosswind force decreases:

(Fig. 6.2): Cross-wind Force and the Over-turning Moment

124
6.2.3 Over-Turning Moment:
These two forces; Drag and Cross-Wind Force appear to
act at approximately the same point on the projectile axis; this
point is known as the center of pressure, as shown in (Fig. 6.2).
As the aerodynamic characteristics of the projectile vary with the
constantly changing velocity and attitude of the projectile, so the
position of the center of pressure also varies.
For all conventional projectiles the center of pressure
does not coincide with the center of mass so these combined
forces produce an over-turning moment dependent on the angle
of yaw. As will be shown later, fin stabilized projectiles are
stabilized by the over-turning moment alone; and without the use
of spin stabilization the overturning moment would cause all other
types of conventional projectiles to tumble.

Finally, the Resistance forces affecting the projectile


during its motion through the medium of the atmosphere can
be represented as in the following (Fig. 6.3):

(Fig. 6.3): Forces affect the projectile motion

125
6.3 METHODS OF STABILIZATION

1) Fin Stabilization
2) Spin Stabilization (Gyroscopes)
3) Fin Stabilization Combined with a very Slow Spinning.

6.3.1 Fin Stabilization (Static Stability):

The problem of ensuring that the projectile will travel


point foremost throughout the flight has been with man since
the earliest of times.

Arrows and javelins were probably the first elongated


projectiles ever used and man soon realized that the best way of
stabilizing the projectile was to concentrate its mass towards the
head. Later tail end fins were added.

(Fig. 6.4): Stability of an arrow

As shown in (Fig. 6.4) G is the center of mass of the arrow. If the arrow yaws and
begins to move obliquely, then owing to the larger surface area towards the rear, the resultant
air pressure acts through the point P, the center of pressure, near the tail and so swings the
arrow round G and onto the trajectory.

126
As long as the center of pressure is behind the center of gravity the projectile
will try to fly nose first. This type of stability is called static stability.

Stabilization by fins is used with aircraft bombs, mortar bombs, as shown in the
following (Fig. 6.5), rockets and Armour Piercing Fin Stabilized Discarding Sabot
projectiles (APFSDS):

(Fig. 6.5): An 81 mm mortar bomb

The fins are fitted to the rear of the projectile to increase the surface area behind the
center of mass. In this way P is brought behind G and the projectile flies in a state of
equilibrium.

Disadvantages of fin stabilization:


Fins, of course, have their disadvantages. They give rise
to large wind effects: it is particularly important in the case
of a rocket being boosted by the burning of the propellant at
the beginning of its flight. The combinations of low-level
wind and motor thrust cause the rocket to turn into the wind.

It is difficult to allow for this adequately in a strong wind, and


it becomes a major source of error in free flight rockets,
which have a long burning time.

6.3.2 Spin Stabilization (Gyroscopes):

127
To withstand the high pressure in the gun, the base of the projectile must be very
strong and therefore thick and heavy. We have also seen that a pointed or streamlined nose is
required to reduce fore-body drag. The combination of the two gives the modern shell its
center of mass near the base, and its center of pressure near the nose. Unfortunately, this
results in an unstable projectile.

Dynamic stability in a shell is obtained by rapid rotation, which produces powerful


gyroscopic forces.

A gyroscope in its simplest form may be considered to be a heavy top. It is a


common experience that a top, if at rest, cannot be made to stand upright, but also the moment
it is spinning fast enough, not only does it stay in an upright position, but also it resists any
efforts to pull it out of that position.

As shown in (Fig. 6.6), the top has a tendency to fall over. It


does not in fact do so, and actually moves around with a sort of
circular motion, that is, it precesses about the vertical axis AB:

(a) Side view (b) Elevated view


(Fig. 6.6): Gyroscopic motion of spinning top

128
If an attempt is made to pull the apex of the top in the
direction shown in (Fig. 6.6.b), it will resist the pull. Instead it will
try to move in the perpendicular direction as shown above.

Precession may be thought of as a circular yaw about the


center of gravity, which takes the shape of a decreasing spiral, as
shown in the following (Fig. 6.7):

Another movement observed with a spinning top is called


nutation. This is a rotational movement in a small circle, which
forms a rosette pattern. Relating these movements to a projectile
in flight we have the effect illustrated in (Fig. 23):

(Fig. 6.7): Precession and nutation for a spinning projectile

Both precession and nutation are rapidly damped out by the gyroscopic action of
the spin and the projectile is stabilized in flight after a relatively short distance.

Now, a top, which is short and squat is much easier to spin, and much less affected by
random disturbances when spinning, than one, which is long and thin.

129
6.4 CRITERION OF STABILITY:

1) Rotating shell:

The criterion of stability for a rotating shell is the same as that for a top, or, in other
words, a shell, which has a large diameter and which is not too long can be made
very stable in flight by the spin imparted to it by the rifling in the barrel.

2) Rockets:

A rocket is usually long and narrow, and to stabilize it by spinning alone is like
trying to make a pencil remain upright by spinning it on its point. It can be done, but
an extremely high rate of spin is needed.

In fact, experiments have shown that, once the length of the projectile exceeds about
“ 7” calibers, a projectile can no longer be satisfactorily stabilized by spinning
alone, no matter what spin rate is adopted. This means that in the case of rockets, fins
provide an easier method of stabilization; although a very slow spin may be
imparted for reasons will be discussed here in after.

130
6.5 OVER AND UNDER SPUN SHELL:
The shell must be made to precess at the correct rate, and keep its nose pointing almost
along the trajectory. The rate at which the shell precesses depends upon the weight
distribution of the shell, its shape, and the position of the center of gravity; but very
important is the rate of spinning of the projectile relative to the rate at which it is moving
forwards and which affects the pressure of the air on it.

Referring to the top again, if it is spun at great speed it precesses only slowly, but as the
rate of spinning dies away the precession becomes more and more rapid.

1) Over-stabilized projectile:

The same effect occurs in the case of a shell's motion about its center of gravity. If the
shell is spun too rapidly it precesses very slowly with the result that the nose never dips
down sufficiently to keep pace with the dipping trajectory and in consequence the shell
lands base first, as shown in (Fig. 6.8). This is the case of the small arms bullet at extreme
range and we say in this case that the shell is over-stabilized.

(Fig. 6.8): Over-stabilized shell

2) Under-stabilized projectile:

Conversely, if the rate of spin is too slow the shell processes too rapidly and the effect
is that the nose dips more quickly than the trajectory. The shell loses considerable
range due to the excessive yaw developed and the consequent tumbling. In this case the
shell is under-stabilized. A loud whirring noise is characteristic of an under-stabilized
shell in flight.

131
CHAPTER 7

PROJECTILE DRIFT

7.1 PROJECTILE DRIFT DEFINITION:


A spun projectile will be found to deviate laterally from the
original direction imparted to it by the gun. This lateral deviation
is called drift as shown in (fig. 7.1):

(Fig. 7.1): The drift of the projectile

In practice, this lateral deviation is made up of two parts,


as will be discussed here in as follows:
1) Drift due to Equilibrium Yaw (for spun projectiles):
Drift due to wind effect, due to gyroscopic effect, and
due to Magnus effect.
2) Drift due to Rotation of the Earth.

132
7.2 CAUSES OF THE JUMP ERROR (Yawing of the projectile):
Previous investigation of jump have shown that it derives
mostly from the barrel bending oscillations due to excitation
by gas pressure during the firing process, the barrel is driven
into high frequency oscillations. The particular oscillation phase
existing when the projectile exits, this is the point in time when
the driving band leaves the muzzle.
Thus, the barrel oscillation has the effect of causing the
projectile to leave the muzzle at an angle, relative to the original
axis of the bore, corresponding to the instantaneous oscillation
phase. In addition, the projectile is given a transverse
velocity, which is proportional to the angular velocity of the
tube.
The bending oscillations of the tube are caused, as already
mentioned, by the internal ballistic processes (pressure of the
propellant gases and velocity curve of the projectile in the barrel),
due to the slight elastic bending in the barrel axis caused by
the gravity and the thermal effect of sun rays during the day
light, or even due to the normal accepted tolerance of barrel
production.
Since usually the center of gravity of the recoiling parts,
and or the trunnion axis lie above or below the axis of the
barrel bore, the gas pressure generates a moment acting on
the recoiling parts, and turning about the trunnion, the
elevation mechanism which can be considered as elastic member
work against this turning moment. Thus, a system capable of
oscillating results, in which the recoil brake, the recuperator
mechanism and the bearings appear as damping members.
These oscillations are now transmitted to the elastic barrel
tube.
For projectiles with eccentric positioning of the center of
gravity, the tube is excited to additional high frequency
oscillations by the centrifugal forces that arise.

133
The coupling conditions between the degrees of freedom of
the recoiling masses, which are assumed to be inherently rigid,
and those of the elastic tube, the elastic properties of the
elevating mechanism, and the resonance frequencies of the
elastic tube, determine the resulting oscillation frequency of
the overall system, and consequently that of the barrel tube.

In addition to tube oscillations, the propellant gases, which


flow out after firing also exert an influence on the projectile
departure; through according to more recent findings, their effects
seem to be of secondary importance.

7.3 DRIFT DUE TO EQUILIBRIUM YAW:

7.3.1 Drift due to Gyroscopic Effect.


In our previous analogy of a spinning top with a spinning projectile, we concluded
that a spinning projectile is subject to gyroscopic forces. In (Fig. 7.2), a pull in the indicated
direction will result in a perpendicular movement as shown below:

(Fig. 7.2): Gyroscopic forces acting on a spinning top

If we now apply the analogy to a shell in flight, we see that


the net result is that; the nose of the shell tends to precess
around the trajectory.

134
Soon after the shell being fired, the shell develops some
vertical yaw with the trajectory. For a shell spinning clockwise
the effect is that the nose will move to the right and the base to
the left, as shown in (Fig. 7.3):

(Fig. 7.3): Rear view of spinning projectile


(Nose movement to the right)

Now, the air pressure will tend to increase the yaw to the
right, but the gyroscopic effect reacts to this air pressure
force, and the nose will tend to move downwards as shown in
(Fig. 7.4):

(Fig. 7.4): Rear view of spinning projectile


(Nose movement downwards)

135
Again, air pressure attempts to increase the nose down
attitude but the gyroscopic effect moves the nose to the left as
shown in (Fig. 7.5):

(Fig. 7.5): Rear view of spinning projectile


(Nose movement to the left)

By continuing in this way, it is easily seen that the net result is that the nose of the
shell tends to precess around the trajectory.

However, whilst this is happening, the trajectory is also dipping. The combination of
a selected clockwise spin rate, precession, and a dipping trajectory keeps the average
yaw more or less constant.

The center of gravity of the shell follows the trajectory and the nose rosettes about
the trajectory with an average position offset to the right. This average yaw to the right is
called "Equilibrium Yaw", as shown in (Fig. 7.6).

Similarly, if an anti-clockwise spin is imparted to the shell we have a drift to the left.

136
(Fig. 7.6): Equilibrium Yaw

7.3.2 Drift due to Wind Effect.


The air stream, which we can imagine coming towards the
shell will therefore, on average, create a high pressure on the
left and push the shell to the right. This is the main cause of
drift. Corrections, which are usually given in Firing Tables, can be
made to compensate for drift.
At very high angles of departure the nose of the projectile
may fail to follow the trajectory at the vertex, with the unfortunate
result that the projectile may fall base first, at a reduced range,
with a marked drift to the left.

7.3.3 Drift due to Magnus Effect.


The Magnus force on a shell results from complex
aerodynamics. However, the fundamentals of the Magnus effect
can be understood by an examination of the effects on a rotating
sphere passing through the air.

137
As shown in (Fig. 7.7), a rotating spherical projectile, for
example an undercut golf ball is traveling through the air:

(Fig. 7.7): Magnus effect

As the ball spins, it will carry some of the air around with it
due to the friction between the air and the surface of the ball; this
effect is increased in the case of a golf ball by the addition of
indentations on the surface.

As there is a meeting of two airstreams travelling in opposite


directions, an increased pressure results at B. It is this increased
pressure, which pushes the ball in the direction from B to A, so
that a resultant lifting force occurs which increases the range of
the ball. At A, we do not have this effect because the two
airstreams are travelling in the same direction. It is the same for
the dipping of a top spun table tennis ball and also for the left
drift of a golf ball when hooked.

138
A shell experiences essentially the same effect. Since, on
average, all shells travel with the nose displaced to the right
of the trajectory, there is a cross flow of airflow from the left
to right of the shell body.

In addition, air is carried around by the shell as it spins. It is


convenient to interpret (Fig. 7.7) as a section through the shell
viewed from the rear. A shell flying with an average equilibrium
yaw to the right will experience an average Magnus force
upwards, which will tend to increase the range. The Magnus
force will also tend to deviate a shell to the left, but the effect is
overwhelmed by the presence of gyroscopic forces, which cause
drift to the right.

Since the Magnus force will not necessarily act through the
center of gravity there may also be a Magnus moment in the
vertical plane. Although the Magnus force is usually small
compared with the equilibrium yaw effect, the Magnus moment
contributes to the dynamic stability and is sensitive to the
spin rate and angle of yaw.

Sometimes the Magnus effect may be large, for example,


when an anti-aircraft gun is fired almost vertically. The yaw at the
top of the trajectory then becomes much greater and so the
Magnus effect becomes more important and the shell may in fact
drift to the left.

139
7.4 DRIFT DUE TO ROTATION OF THE EARTH:

When dealing with long ranges the effects due to rotation of


the earth must be taken into account. The rotational speed of the
earth's surface at the equator is about 450 m/s, i.e. 1620 km/h. At
first sight this appears to be of no consequence as the gun, target
and atmosphere are all travelling together, more or less, at the
same speed.
However, the motion of the earth's surface is, not in a
straight line, so during the period that a projectile is in flight
the target may be carried sideways, up or down by the earth
to which it is attached.

Fortunately the rotation rate and dimensions of the earth are


fixed, so for any gun and target at known positions on the earth's
surface it is possible to apply corrections to the gun bearing
and elevation to ensure that the projectile meets the target.

A rigorous treatment of this drift is beyond the scope of this


book. However, it is not difficult to get an adequate grasp of the
principles involved.

As the earth's rate of rotation is constant, the angle through


which the earth rotates during a projectile's flight is dependent
upon the time of flight, and the range of the trajectory magnifies
the extent of the resultant drift. Consequently, drift of this form is
of significance only to trajectories featuring both long range
and long time of flight.

Of course long ranges imply long flight times, so very long


ranging projectiles, notably ballistic missiles, are severely
affected by this drift.

140
The drift of artillery shells is much less. At a range of 20 km
the drift may be 100 m, depending on the gun's geographic
position, the bearing to the target and the time of flight. Over
ranges of less than 5 km the drift is generally much less than the
shot to shot variations of point of impact, whilst for small arms
the drift due to rotation of the earth is so small that it is
always ignored.

7.4.1 Types of drift due to rotation of the earth:


There are three forms of drift that can be distinguished.
The first two drifts primarily affect the range of the projectile, and
the third drift mainly affects the bearing to the target:
1) The first form of drift:
It is observed for trajectories between points at the same
latitude. When firing east, in the same direction as the rotation of
the earth, the earth effectively rotates down and away from
the trajectory so the projectile passes over the target and strikes
the earth some distance beyond, as shown in (Fig. 7.8):

(Fig. 7.8): Extension of range


(When firing eastward at low elevation)

141
Conversely, firing west results in the projectile falling short
of the target. Whether firing east or west, this form of drift is
always eastward. In addition, this same effect causes a very
slight sideways drift of trajectories at moderate latitudes as the
projectile follows a great circle route while the target rotates
equilaterally.
2) The second form of drift:
The second form is most apparent when a projectile is
fired directly up wards at the equator. During its period of flight
the earth rotates so that the projectile lands at a point west of the
launching point.

3) The third form of drift:


For the final form of drift consider a gun firing across
the North Pole. During the time of flight the earth rotates anti-
clockwise beneath the trajectory, so causing the projectile to land
to the right of the point of aim. This drift to the right occurs
throughout the northern hemisphere, though lessening towards
the equator and absent at the equator itself. The effect is reversed
in the southern hemisphere, so causing a drift to the left.

The first two forms of drift are greatest at the equator and
lessen towards the poles. They have an opposing effect upon
range; the first dominates somewhat greater drift at lower
elevation trajectories, and the second dominates drift at very high
elevations of the gun. At a certain elevation these two drifts are
equal and opposite, so they have no overall effect upon the
range, this elevation is generally about 60°.

These three forms of drift have a superimposed effect. For


guns, the required adjustments to bearing and elevation to correct
for the drift are tabulated in firing tables to suit all possible
trajectories.

142
CHAPTER 8

CHARACTERISTICS OF A FREE FLIGHT ROCKETS

8.1 VELOCITY CONSIDERATIONS:

In this section we will look at the basic characteristics of a free


flight rocket; that is, a rocket which has all its guidance given to it
by the launcher.

A rocket consists of many parts. The head is designed as a


shell, while the rear part of a rocket is normally taken up by the
motor, fuel and combustion chamber.

The combustion chamber is analogous to the chamber of a


gun in which the propellant is burned. By means of a nozzle
system the products of combustion are ejected at high speed
from the rear of the motor. These gases exhaust with a certain
momentum, and according to Newton's Second Law of Motion,
the rate of change of momentum of the gas leaving the rocket
represents a force, and, according to Newton's third law, this
force is reacted against by a thrust in the opposite direction.
It is the thrust that drives the rocket forward.

Rockets may have either liquid or solid fuels. Generally


speaking liquid fuel rockets have a high energy content, while
the solid fuel rockets are relatively simple to produce and easy
to handle, but have a relatively low energy output.

In a vacuum, the maximum velocity of a rocket depends on


the speed of the emerging gases and on the ratio of the weight of
the rocket before firing to the weight when the propellant is all
burnt. This ratio is called the "mass-ratio".

143
The acceleration at any moment depends on the rate at
which the fuel is being consumed, the speed of the emerging
gases, and on the remaining weight of the rocket. The maximum
speed will be reached just when the fuel is completely burnt.

When the rocket is fired in the atmosphere these statements


are only approximately true. All the external effects must be
considered, as was the case with shells.

In general, the range of a rocket may be increased by


choosing the composition of the fuel so that the speed of the
emerging gases, or exhaust velocity, is as large as possible, and
by designing the projectile to make the mass-ratio as large as
possible.

8.2 ROCKET ACCURACY:

The principal disadvantage of free flight rockets is their


relatively poor accuracy and consistency compared with guns.
The main sources of error are due to wind effects, thrust
misalignment and velocity at all burnt.

As the rocket is moving slowly when it leaves the launcher,


surface side wind effects on the stabilizing fins can be large, and
the rocket tends to turn into the wind. Wind corrections are
therefore applied at the latest possible moment before firing to
reduce this error. The thrust from the rocket motor must act along
the axis of the rocket. If it is offset it will tend to act like a rudder
and deflect the rocket off course. This misalignment can be
partially overcome by spinning the rocket slowly. A small change
in the rate of burning of the rocket propellant can make a
significant difference to the velocity of the rocket at all burnt, which
in turn will affect the final range achieved. Errors can also arise if
the rocket center of gravity is not on the nozzle axis.

144
8.3 ROCKET STABILITY:

There are three main methods of achieving stabilization of free flight rockets:

1. Fin stabilization

2. Spin stabilization

3. Combination of fin and spin stabilization.

Large fins are required to minimize the flight deviations arising from launch and thrust
misalignment, but small fins are necessary to reduce the deviations arising from surface cross
winds. The optimum size of fin is chosen to achieve the least overall dispersion.

Slow spin is imparted during launch and the resultant motion of the free flight rocket
aids the rocket in maintaining its stability about its vertical, longitudinal and lateral axes.

Most free flight rockets achieve stability by spin and fin stabilization. Three principal
methods used to impart slow spin are:

 A slow rate of spin imparted, by special arrangement in the launcher, to the rocket
during its motion on the launcher.

 A slow rate of spin imparted, by using an inclined nozzle system with respect to the
rocket axis, to the rocket during its flight.

 Auxiliary rocket motors, which are fired to spin the rocket immediately after it is
launched.

145
8.4 ROCKET TRAJECTORIES:

So far we have been considering trajectories of


projectiles such as cannon shells, mortar bombs and
projectiles, which are not powered in flight. Let us now
consider the trajectories of rocket-propelled projectiles,
which are powered. The motion of a rocket may be divided
into three phases:
 The launch phase
 The boost phase
 The ballistic phase
In launch phase the launch tube is used to direct the
initial motion of the rocket. Launch dynamics is a complete
study in itself and may be discussed in another course.

Once the rocket has accelerated to its maximum velocity


we have an "all-burnt" situation and the motion enters the
ballistic phase. This part of the trajectory is similar to that of a
shell and will not be discussed further.

The period of greatest concern for rocket accuracy is the


boost phase between launch and all-burnt, during which time,
under the influence of rocket thrust, the motor is sensitive to
initial disturbances arising from the launch: two of these are
wind and thrust misalignment. A full theoretical account of
rocket motion may be found in Rankin.

In the simplest treatment of the rocket trajectory it is

146
assumed that launch is perfect and the rocket moves under
the influence of gravity, axial thrust and drag forces only. We
can therefore assume that the rocket behaves as a point
mass. We ignore all the forces and couples acting on the
projectile, as these will, on average, often cancel each other
out.

The point mass assumption can be used in the


preparation of range tables and ballistic data.

8.4.1 Rocket Thrust:

The action of a rocket motor depends on conservation of


linear momentum. Let us consider the situation as depicted in
(Fig. 8.1). The total momentum at time t is equal to the total
momentum at time t + δt:

Time: t Time: t + δt
Mass: m V* ◄▬ δm ◄▬ Mass: (m – δm)
▬► ▬►▬►
Velocity: V + δV ▬►
Velocity: V ▬►

(Fig. 8.1): Conservation of linear momentum

In the short time interval at a particle of mass δm is ejected


with a velocity V* relative to the rocket. Equating the momentum
before and after the ejection of this particle:

m V = (m – δm) (V + δV) + δm (V + δV – V*)

then by solving this last equation, we obtain:

m δV = δm V* = - (dm/dt).δt . V*

147
where: δm/δt = - dm/dt i.e. δm = - (dm/dt).δt and dm/dt: is the
rate of change of mass that is negative since m is in fact
decreasing with time.

dividing by δt, we get:

m (δV/δt) = - (dm/dt) V*

and by letting δt→ 0, we obtain the differential equation of


motion:

m (dV/dt) = - (dm/dt) V*

where:
V*: is known as the exhaust velocity of the rocket and
may quite often be taken as a constant.

The expression T = - (dm/dt) V* may be regarded as the


thrust of the rocket. In this case the equation of motion can be
integrated to give:
Vo  dV = - V* mo  m dm/m
V

V - Vo = V* loge mo/m

where: Vo: is the initial speed of the rocket


mo: is the initial mass.

In the above calculation air resistance has been neglected.


The thrust, [ T = - (dm / dt) V* ] is quite often very large compared
with the drag and gravity so that this omission is justified. The
equation of motion neglecting gravity but including drag effect is:
m (dv / dt) = T - D
where:
T : is the thrust of the rocket
D: is the drag.

148
In practice V* is the effective exhaust velocity since the
thrust of a rocket is composed of two parts:
 Mass loss term.
 Pressure term.
Here the thrust is given by:

T = - dm / dt . [ V* ]
= - dm / dt . [ V′ ] + (Pe – Pa) Ae
where:
V′ : is the true exhaust velocity, [ V′ = V*- (V + δV)]
V*: is the effective exhaust velocity.
Ae: is the exit area
Pe: is the exit pressure
Pa : is the atmospheric pressure
Additionally the quantity V*/g which has the dimension
seconds is called the specific impulse.

Specific impulse depends on the propellant: for the “V2-


rocket” it was 205 (s), for the "Redstone" missile 235 (s) and for
the "Atlas" missile 250 (s).

(Table 8.1) shows some exhaust velocities and specific


impulses for various types of propellant:

Specific Impulse Exhaust Velocity


Types of propellant (sec) (m/sec)

75% ethyl alcohol


And li quid oxygen 279 2733

Liquid hydrogen
And liquid oxygen 391 3837

149
Fluorine
And hydrogen 410 4024

95% hydrogen peroxide


And RIP 273 2680

(Table 8.1): Some exhaust velocities and specific impulses


for various types of propellant

It is also important that the mass ratio “mo/m” (the mass at


launch over the mass at burnout) is high; for a "V2" rocket the
mass ratio was 2.8, whilst a modern rocket has a typical value of
about 5.

150
8.4.2 Rocket Drag:

The drag whilst the motor is burning is different from


the free flight drag and so adjustments are made to account for
base drag. The situation is thus taken to be:

Drag during burning = Total resistance of shell of same caliber


and head shape (nose drag, base drag
and skin friction)
+ Skin friction of rocket
- Skin friction of shell
- Base drag over the nozzle area of
the rocket.

For conventional artillery rockets the exhaust plume removes the base drag con-
tribution during boost, so:

Drag after all-burnt = Same as above but omitting base drag over
nozzle area term.

The skin friction drag is taken as ½ ρ V2 S CF


Where:
S : is the surface area of the rocket
CF: is the skin friction drag coefficient.

Usually turbulent flow is assumed fort which:

CF = 0.455 (log10 Re)-2.58


where:
Re = V L/
L: is a typical dimension of the rocket body.

The skin friction acting on fins is usually considered


separately, L being their stream wise chord length.

Recently, results of wind-tunnel tests have been used to


calculate drag on a variety of rocket-propelled projectiles,

151
8.4.3 Equation for Rocket Trajectories:
The equations of motion along and at right angles to the trajectory as shown in (Fig.
10.2) may be written as:

(Fig. 10.2): Rocket trajectories

m. dV/dt = - m• . V* – R – m g sin θ

m .V. dθ/dt = – m g cos θ

If (- m• V*) is large compared with (R) and (m g sin θ) we


may integrate the first equation to give V at all burnt:

V = V*. loge (mo / m)


The acceleration (dV/dt) is often uniform so that along the trajectory the velocity V
and the distance X will be:

V = a.t

X = ½ a.t2

where:
a: is the uniform acceleration.
The angle θ may then be found by putting (V = a.t) into the second equation:

152
d θ = (g. cos θ)/ a. dt/t .

consider (cos θ) to be constant and equal to (cos θo), and integrating from θ to θo and from t to
to, to obtain:

θ = θo – (g / a) cos θo . loge (t / to)


here (to) is the instant the rocket leaves the launcher with
velocity (V = a.to).

If however (- m• V*) is not large compared with drag etc.


and wind terms are important we have the following
equations of motion:

a) During burning:
m v • = [- ½ ρA lv – wl2 ĈD - ue (dm/dt) ][ (v – w)/ lv – wl ] – m g ĵ
where:
v : velocity = ux î + uy ĵ + uz ĸˆ ,(y: is now vertical coordinate)
W: Wind velocity= wx i+ wz k ,(i.e has no vertical component)
dm / dt = Rate of loss of mass
ue= effective exhaust velocity
V*= specific impulse x grafity
ĈD= drag - (base drag)

b) After all burnt:

m v • = [½ ρA lv – wl2 CD][ (v – w)/ lv – wl ] – m g ĵ


where:
CD: is the drag plus base drag.

A computer program may be written to solve the above equations.

153
CHAPTER 9

TRAJECTORIES

9.1 CONCEPT:

In this section of external


ballistics a deep and more theoretical study of some of the basic
concepts will be introduced.

It is now necessary to refine our views by examining in more


detail the various aspects such as trajectories, fluid motion and
aerodynamics. A study of these areas of ballistics is vital in order
to gain a complete understanding of the in-flight behaviour of
projectiles.

9.2 TRAJECTORIES:

9.2.1 Assumptions:

A large part of external ballistics consists of evaluating


trajectories. There are numerous forms of trajectory modelling
involving different basic assumptions and having different
complexities of solution.

The majority of trajectory modelling undertaken nowadays


uses numerical integration by computer and this has had the
effect of rendering obsolete many of the analytical and semi-
empirical techniques.

In this section we shall assume that the projectile behaves


as a point mass and only consider the influence of gravity, drag
and wind effects. Once we leave point mass trajectories, other
considerations such as the angle of yaw, aerodynamic
moments, etc, have to be made; and a treatment of these effects
in trajectory modelling is beyond the scope of this chapter.

154
155
9.2.2 Computation of Trajectories by Computer:

The methods of calculating trajectories using tables have


been largely superseded by the use of computers, such as
Prodas package, which integrate the x, and z differential
equations of the trajectory illustrated in (Fig. 9.1). These are, in
aerodynamic notation:

(Fig. 9.1): The effect of gravity and drag acting on a projectile

m x▪▪ = - ½ ρ V2 A CD cos θ°
= - ½ ρ V2 A CD (x▪/ V)
m z▪▪ = -½ ρ V2 A CD sin θ° - mg
= -½ ρ V2 A CD (z▪/ V) - mg
where:
V2 = x▪2 + z▪2
with initial conditions when:
t = 0:
xo = 0, zo = 0, x▪ = Vo cos θ, z▪ = Vo sin θ

156
The above equations will be integrated by a step-by-step
process.

They are first rewritten as a set of first order differential


equations by introducing the new variables:

Vx = x ▪
And:
Vz = z ▪

The first order equations may then be integrated by a fifth


order Runge-Kutta technique. A variety of instructions can be
keyed in and a graphical output obtained.

157
CHAPTER 10

FLUID MOTION

10.1 FLUIDS:

The word fluid is used to describe both liquids and gases. Of


particular interest to ballisticians is the resistance of a projectile to
motion caused by the viscosity or friction of the air.
We have already defined the absolute viscosity () of a gas
as a measure of the flow resistance of the gas, and the
kinematics viscosity ( =  / ) as the ratio of absolute viscosity
to density.
We now define another important quantity called the
Reynolds Number (Re):
Re = Inertia force / Viscous force
The inertial force acting on a typical fluid particle is
measured by the product of its mass and its acceleration.

Now the mass per unit volume of the fluid is, by


definition, the density , while the volume of the particle is
proportional to the third power of the length scale. Hence the
mass is proportional to L3, or simply is proportional to L3.

The mean acceleration is the change of velocity divided by


the time in which this change occurs.
The changes in velocity as the fluid accelerates and
decelerates in the neighbourhood of the body are of the same
order of magnitude as the speed, V, itself. The time in which this
change occurs is of the same order as the time for an average

158
fluid particle to travel the length L of the body at the speed V. This
time is (L / V).

159
Accordingly the acceleration is proportional to the velocity V
divided by (L / V), that is:
acceleration  (V2 / L)
Therefore:
inertia force = mass x acceleration
~ L3. V2 / L
= L2 V2
where:
~ : means both
"proportional to" and "of the same order as".

The viscous forces are determined by the viscous shear


stress acting on the body and the surface area (L2) over which it
acts. The viscous stress is proportional to viscosity () and to the
rate of change of speed with distance, that is (V / L). Therefore:
Viscous Forces =  L2. V / L
=LV
We can now define Reynolds number more precisely as:
Re = Inertia force / Viscous force
~  L2 V2 /  L V
= LV / 
= LV / 
where:
V: is velocity (m/sec)
L: is length (m)
: is absolute viscosity (kg/ m.sec)
: is kinematics viscosity (m2/ sec)

We are now in a better position to define the drag coefficient.


160
10.2 DRAG COEFFICIENT:

The drag force D acting on a projectile in flight is usually


written in the form:

D = ½  V2 A CD
Where:
 : is the air density ( kg / m3)
V : is the velocity of projectile (m / sec)
A : is the reference area of projectile (m2)
CD: is the drag coefficient

CD is dimensionless and is a function of two further


dimensionless quantities:
Reynolds number, Re = L V / .
Mach number, M = V/a

Where:
L: is a typical dimension
a: the velocity of sound in air.

Reynolds number represents the effect of viscosity and


Mach number represents the effects of compressibility.
In the SI system of units:
The density of dry air at 15°C and a pressure of one atmosphere
[1.013 x 105 (N/m2)] is equal to 1. 225 (kg/m3).
The viscosity of dry air is 1.78 x 10-5 (kg/m.sec) and the kinematic
viscosity ( =  / ) = 1.45 x 10-5 (m2/sec).
The velocity of sound in air under these same conditions is 340.6
(m/sec)

In addition to the components described above, CD depends


on the shape of the projectile and the yaw angle.

161
10.3 DYNAMIC SIMILARITY:
If the Reynolds number Re is the same for similarly shaped
objects in different fluids, then the flows about the objects are
equivalent. This equivalence is called dynamic similarity. Data
provided by Shapiro shows exactly how the principle of dynamic
similarity makes it possible to predict the performance of full-scale
Projectiles (or aircraft) from relatively convenient and
inexpensive wind tunnel tests of scale models.
Visc Reyn
Length Speed Fluid Density Drag Drag
osity old’s
Objects Kind Number Coefficient Force

L V   Re CD D
Helium-filled
100 5.4 Air 0.0012 0.018 36 0.000019 0.0067
Balloon
Plasti
0.60 55 Water 1. 00 0.89 37 0.000018 0.020
c Ball
(Table 10.1): Dynamic similarity

For comparison a small plastic ball with a diameter of 0.60


(cm) was dropped into water, while a large helium-filled balloon
with a diameter of 100 (cm) was allowed to rise in air. For the
small ball the Reynold's number was calculated to be 37, and for
the gas balloon was 36. Thus, for these two particular
experiments the Reynold's number differs very slightly. So while
these experiments, one large ascending in gas, and the other
small descending in a liquid, one, appear very different to the eye,
they are in fact dynamically similar. The important practical
implication of this is that the condition of dynamic similarity
enables us to predict the drag of the helium balloon in air, by
making a different experiment in which we measure the drag of
the plastic ball in water.
Dynamic similarity is the key to how we are able to
predict, from the measurements made in the wind tunnel on a

162
model projectile (such as a rocket), the forces which the real
projectile (rocket) would actually experience.

10.4 THE RELATIONSHIP BETWEEN THE DRAG OEFFICIENT


“CD” AND REYNOLD’S NUNBER “Re”:

It may be shown that when experiments are carried out on


geometrically similar shapes at the same Reynold's number, the
drag forces on the objects, while perhaps very different in
magnitude, are related in a very special way.
Referring to the experiment with the ball and balloon we
again see that the drag coefficients are very similar, as shown in
the above table. This leads to the law of dynamic similarity:
equality of Reynold’s number implies equality of drag
coefficient.
A typical graph of the variation of the drag coefficient with
Reynold's number is given below as shown in (Fig. 10.1):

(Fig. 10.1): The relationship between CD and Re

163
10.5 LAMINER AND TURBULANT FLOW:

Laminar flow in a fluid is easily seen in the smooth


streamlined motion observed in freely flowing thick oil. It is
difficult to make the oil turbulent or chaotic by mixing since any
irregularities in the flow are quickly damped out by strong viscous
action. In laminar flow the fluid particles move downstream in
smooth and regular trajectories, without appreciable mixing
between different layers of fluid. Such a laminar flow is termed
steady.

In a turbulent flow there is superimposed on the average


downstream motion of the fluid an irregular and seemingly random
motion. In the mixing process any slow-moving fluid is invigorated
and speeded up by fast-moving fluid with which it mixes, while the
latter in turn tends to be slowed down.

We can determine whether a fluid motion will be laminar or


turbulent by considering the size of Reynold’s number.

If Re is very large, say of the order 107, we know that the


inertia forces will dominate, and if Re is very small, say of the
order 102, viscous forces will dominate. In general, if:

Re < 104, the flow is orderly and steady and this


is defined as “laminar flow".

Re > 106, the flow is very energetic and this is


called "turbulent flow". '

164
10.6 VISCOUS BOUNDARY LAYER:

In all fluids the particles immediately in contact with solid


boundaries do not slip relative to the boundary. Therefore, while
most of the air may rush past a projectile at high speed, the air at
the surface of the projectile does not move at all relative to the
projectile. Between this film of relatively motionless air next to the
projectile and the fast moving main body of air at some small
distance from the projectile, there is a region of flow called the
boundary layer, as shown in (Fig. 10.2):

Non-viscous flow

Boundary layer

(Fig. 10.2): Viscous boundary layer

In this layer the speed changes from zero at the surface, to a


velocity virtually unimpeded by viscous forces at the outside edge
of the layer. The formation of the boundary layer is initiated by the
motionless layer of fluid at the surface of the projectile. This
motionless layer exerts a viscous drag on the layer of fluid next to
it, thus gradually slowing down that second layer. This effect
continues from layer to layer. In this way the decrease of velocity
at the surface is diffused by viscosity outward through the fluid,

165
and more and more fluid becomes caught up in the viscous
boundary layer.

The continual thickening of the boundary layer is shown in


general for a fluid stream travelling from left to right past a flat
surface. The arrows represent the thickness of the layer at
successive instants of time.
The amount by which the boundary layer grows in a certain
time or in a certain distance depends on such things as the
speed, the viscosity, and the density.
In general, as the Reynold's number increases, the
boundary layer grows less rapidly. It should be noted however;
that no matter how large the Reynold's number becomes the
viscous boundary layer can never totally disappear.
As Re decreases, the boundary layer becomes thicker and
thicker. At extremely low Re, the viscous region extends very far
from the body, and occupies practically the entire region of flow. It
is meaningless then to even speak of a boundary layer.
In general the boundary layer grows in thickness from the nose
of a projectile, and has a different law of growth according to the
type of flow.

 For a laminar boundary layer, the growth is given by:

lam / x = 5.5 Rex-0.5


where:
lam: is the thickness at a distance x from the nose or
leading edge
Rex: is the Reynold's number at the same point.

 For a turbulent boundary layer, the growth is different


and is given by:

166
turb / x = 0.22 Rex-0.2

For ballistic purposes we are mostly concerned with


turbulent flow and turbulent boundary layers. The state of the
boundary layer indicates:

1. The effect that protuberances or roughness have had on


the skin friction.

2. The sensitivity of the flow field to body curvature.

3. The extent of the cavity in the base region of the body.

Two ballistic examples are:

1. For a 7.62 mm caliber, 28 mm long bullet, having a


muzzle velocity of 840 m/s:

Re = L V / 
= (0.028 x 840) / (1.45 x 10-5) = 1.62 x 106

Where the kinematic viscosity of air is taken at sea level.


A turbulent boundary layer is indicated with a boundary
layer thickness of 0.35 mm at the bullet base.

2. For a 9 m long rocket travelling at velocity 600 m/s at an


altitude of 5.6 k:

Re = L V / 

= (9x 600) / (1.45 x 1.58 x 10-5) = 2.36 x 108

In the second example we have used a formula relating


the kinematic viscosity at sea level and some other height.
The Reynold's number found represents a well

167
established turbulent boundary layer with a thickness at
the rocket base of 42 mm.

At high Re there is separation of the airflow while at low Re


there is not. At first sight it may seem natural to think that a
turbulent boundary layer would produce a larger drag force.
However this is not necessarily true, because a turbulent
boundary layer remains attached longer and produces a smaller
wake.
This effect is most readily observed when the flight of a golf
ball is considered. The boundary layer becomes turbulent for a
roughened sphere earlier than it does for a smooth sphere and
consequently a smaller wake is produced due to the boundary
layer separation, as shown in (Fig. 10.3):

(Fig. 10.3): Boundary layer separation


Although the skin friction is increased for a roughened
surface, the reduced wake causes a greater drop in base drag
and a subsequent drop in total drag. At some speeds the drag on

168
the roughened sphere is only about I/5 of the drag on a smooth
sphere. This is the main reason why golf balls are dimpled.

169
CHAPTER 11

AERODYNAMICS

To achieve a complete understanding of the flight of a


projectile it is essential that all the forces, moments and other
terms affecting the flight of the missile to be represented in well-
defined mathematical form.

11.1 AERODYNAMIC FORCES AND MOMENTS ACTING ON A


PROJECTILE:

There are two distinct types of major forces and moments,


which act on a projectile in flight. They are:
1) Static Forces and Moments:
(Drag force, Over-turning moment):
Which are dependent on the attitude and translation
velocity of the missile?
2) Dynamic Forces and Moments:
(Damping force, Magnus force, Magnus moment, Spin
damping moment, Damping moment):
Which are dependent also on the angular velocity.

We shall not consider in detail the Magnus force or


damping force, during this course.

Now, when the projectile has zero yaw the only force acting
against the projectile motion is the drag force but when the body
has an angle of yaw to the relative wind then we have a net
cross wind force: this is sometimes called the lift force.

170
Consequently the drag and lift force are in line with and
perpendicular to the trajectory respectively. These are known as
"wind -axes", as shown in the following (Fig. 11.1):

(Fig. 11.1): Wind-axes

The American literature commonly refers to the axial force


and normal force, which are in line with and perpendicular to the
body axis respectively as "body-axes", as shown in (Fig. 12.1):

(Fig. 11.2): Body-axes

Since in ballistics we are usually concerned with the motion


of the center of gravity of the projectile along the trajectory we
shall use wind-axes.

171
On every axi-symmetric body shape there is a point on the
axis where the resultant force appears to act, about which
therefore the net moment is zero. This is the center of pressure,
and for incompressible flow it is a function of geometry only; but in
compressible flow and especially at transonic speeds the position
of the center of pressure is sensitive to Mach number and
changes as the Mach number changes.

Because the center of gravity of the body is the point that


defines the trajectory it is convenient to make our forces act at the
center of gravity. We can transform the resultant force at the
center of pressure to a couple together with a force at the center
of gravity. We then have the drag and cross wind force acting at
the center of gravity, and a moment is called the over-turning
moment for a shell, which is usually aerodynamically unstable,
and the restoring moment for a missile, which is usually
aerodynamically stable. We shall call it the yawing moment M
and define it as positive with increasing angle of yaw .
It is convenient to use non-dimensional coefficients for
describing the forces and moments acting on the projectile. It is
usual to non-dimension alise a force by dividing by (½ V2AO) and
a moment by (½ V2AO.d), where: d is the caliber of the projectile.
Thus we have:

Drag coefficient:
CD = Drag Force / (½ V2AO)
Cross Wind coefficient:
CL = Cross Wind Force / (½ V2AO)
Yawing Moment coefficient:
CM = Yawing Moment / (½ V2AO.d)

172
11.2 VARIATION OF AERODYNAMIC COEFFIIENTS WITH
ANGLE OF YAW:

11.2.1 Drag Coefficient CD Versus Angle of Yaw :


With an axi-symmetric body the variation of CD is
symmetric with , and has a minimum value for zero yaw. As
shown in (Fig. 11.3) we can write:
CD = CDo + CD2. 2
where:
The term: CD2. 2 is sometimes called the yaw-drag
coefficient and represents the increment in drag coefficient
resulting from the angle of yaw.

(Fig. 11.3): Drag Coefficient CD Versus Yaw Angle 

11.2.2 Cross Wind Coefficient CL Versus Angle of Yaw :


For angles of yaw less than about 10° projectile nose
shapes give rise to lift forces proportional to the angle of yaw.
Since an axi-symmetric body has zero lift for zero angle of yaw
the function is a straight line passing through the origin.

173
From the equation of a straight line as shown in (Fig. 11.4)
we can write:
CL = (∂ CL/∂ ) 
That is:
CL = CL . 

where the sub-suffix denotes differentiation with respect to .

(Fig. 11.4): Cross Wind Coefficient CL Versus Yaw Angle 

11.2.3 Yawing Moment Coefficient CM Versus Angle of Yaw :

Disturbances, such as muzzle motion during projectile exit


or cross wind gusts will impart a yawing moment. The yawing
moment causes yaw acceleration and the angle of yaw will
continue to change even though the disturbance has ceased.

If the yaw angle continues to increase, the moment is


destabilizing and the projectile is aerodynamically unstable. We
express this as:

∂ CM / ∂ = CM  > 0

174
The center of pressure of such a body is located
between the nose and the center of gravity. On the other hand,
if the derivative CM  < 0, the body is statically stable and the
center of pressure is located behind the center of gravity.

In this case the generated yawing moment strives to


decrease the absolute value of the yaw angle and to return the
body to its initial position. A damped oscillation commonly
occurs.

If the derivative CM  = 0, the body is neutrally stable and


the center of pressure coincides with the center of gravity. This
property is often used in wind tunnels and ballistic ranges
where the center of gravity of the model is adjusted until the
yawing moment is zero.

(Fig. 11.5): Yawing Moment Coefficient CM Versus Yaw Angle 

Knowledge of the aerodynamic characteristics of a


projectile is essential for the designer.

175
There are many ways of obtaining estimates of the relevant
aerodynamic coefficients. The main ways are:

1. Theoretical estimation using fluid dynamic theory


2. Empirical estimation using data on similar projectile
shapes
3. Wind tunnel testing
4. Range testing
5. Firing trials data, including Yaw Sonde and Radio
Doppler methods.

For most projectile designs almost all of these techniques


are used before the design is completed.

176
11.3 GYROSCOPIC STABILITY:

For reasons have given before herein, conventional shells


are invariably aerodynamically unstable. Nevertheless it is
essential that, after leaving the gun, projectiles should continue to
point in the general direction of their flight; otherwise drag forces
would be unacceptably high and variable side forces would make
the trajectory unpredictable and precision would suffer. Flight
stabilization is achieved by spinning the projectile. To see how
stability is actually attained it is informative to solve the equations
of motion for a spinning top first.

11.3.1 Spinning Top:


We consider a symmetrical top, that is, a rigid body that is
a solid of revolution. Let one point on the axis of symmetry is fixed
at the origin 0 of coordinates. We then have the situation shown in
the following (Fig. 11.6):

177
(Fig. 11.6): Spinning top
Let:
 The mass of the top be m.
 The Moment of Inertia about axis of symmetry “OC” be A.
 The Moment of Inertia about axis perpendicular to “OC”
through”0” be B.
The forces acting on the top are:
1. Weight = mg.
2. Reactions at the fixed point O.
We will ignore any frictional couple at O. Then let the top
spins with angular velocity ω relative to plane ZOC, i.e. about axis
OC. Motion due to change of u is called Nutation, motion due to
change of v is called Precession. The general motion of the top is
one of nutation and precession.

 The Angular Velocities:


v• about OZ, u• about line perpendicular to OH in plane XOY,
i.e. OT and OD about OC (relative to plane ZOC), as shown in
(Fig. 11.7):

178
(Fig. 11.7): Angular velocities of spinning top

Resolve v along axes OC, OD where OD is in the plane COH
and perpendicular to OC, then:
v• cos u along OC
v• sin u along OD
Thus angular velocities are:
ω + v• cos u about OC
u• about OT
v• sin u about CD
The corresponding angular moments are:
A (ω + v• cos u) about OC
B u• about OT (1)
B v• sin u about CD
Let: N = ω + v• cos u
(2)
Then: N is the rate of spin or total angular velocity about OC.
The components along OC and OD are equivalent to:
A N cos u + B v• sin2 u about OZ
A N sin u - B v• sin u .cos u about OH
Finally resolve the components about OH, OT into components
about OX, OY, as shown in (Fig. 12.8) to give:
about OX:
Hx = (A N sin u - B v• sin u .cos u) cos v - B u• sin v
about OY: (3)
Hy = (A N sin u - B v• sin u .cos u) sin v + B u• cos v

179
about OZ:
Hz = A N cos u + B v• sin2 u

(Fig. 11.8): Resolved Components of Angular Velocities

We now have the angular moments measured from fixed axes


OX, OY, OZ. Hence we have the equation of motion as (on taking
moments about axes: OX, OY, OZ):
dHx /dt = - m g L sin u. sin v
dHy /dt = m g L sin u. cos v
dHz /dt = 0
where:
L = OG
Equations for small oscillations, i.e. u is assumed small If u is
small then we have:
sin u  u
cos u  1
and so from (3)we have:
Hx = A N u cos v – B (v• u cos v + u• sin v)

180
Hy = A N u sin v - B (v• u sin v - v• cos v) (5)
Hz = A N

since terms of (θ2) are neglected, Let:


x = sin u. cos v  u cos v
y = sin u. sin v  u sin v
i.e. x, y are the projection of a unit length on OC onto plane XOY.
d/dt (A Nx - B y•) = - m g Ly

d/dt (A Ny + B x•) = m g Lx

d/dt (A N) =0
Thus:
N = constant
Writing:
Q = A N/ B
F = m g L/ B
we have:
d/ dt (Qx - y•) = - Fy
d/ dt (Qy - x•) = Fx (7)
If we put:
z = x + yi
then:
z•• - i Qz• = Fz (8)
if we let:
z = W e1/2 iQt
then (8) becomes:
W•• + (1/4 Q2 – F) W = 0 (9)

181
equation (9) represents simple harmonic motion provided:
2
1/4Q – F > 0
that is, provided the stability coefficient, s, is such that:
s = Q2/ (4F) > 1
Now:
s = Q2/ (4F) = A2 N2/ (4 B2 F) = A2 N2/ (4 B mg L)
so that a minimum spin rate which is necessary for stability will be:
N = (4 B mg L)1/2/ A

11.3.2 General Stable Motion:


Note that we view the top from above. Assume stable
motion, that is:
S>1
then (9) has solution:

W = R exp { i[-(1/4 Q2- F)1/2 t + e]} + S exp { i[(1/4 Q2- F)1/2 t + f]}
(10)
Where:
R, S: are positive real arbitrary constants
e , f : are real arbitrary constants.

We can interpret the motion as follows:


 The first term in equation (10) represents an arm OA of
length R inclined at an angle f initially and rotating
clockwise with rate (1/4 Q2- F)1/2.
 The second term represents an arm S (AB) initially
inclined at an angle e to x-axis and rotating with angular
speed (1/4 Q2- F)1/2 in an anti-clockwise sense.
As shown in the following (Fig. 11.9), the combined motion
is that of point B where the arms are joined at A. In general B
182
describes an ellipse in the W-plane with major axis (R+S) and
minor axis IR- sl. The period is [2 π/ (1/4 Q2- F)1/2].

(Fig. 11.9): Two arm model applied to a spinning top

From definition of W and equation (10) we have:


W = R exp{ i [(1/2 Q - (1/4 Q2- F)1/2) t + e]}

+ S exp{ i [(1/2 Q+ (1/4 Q2- F)1/2) t + f]} (10)

In the z-plane the angular rate of rotation of the arms are


(in an anti-clock-wise sense) as follows:

 R arm: 1/2 Q - (1/4 Q2- F)1/2 slow.


 S arm: 1/2 Q + (1/4 Q2- F)1/2 fast.

183
CHAPTER 12

THE SPINNING PROJECTILE

12.1 FORCES ACTING ON A SHELL:


In addition to gravity
there will be aerodynamic forces acting on the shell, as already
discussed. All of the additional forces can be reduced to a single
force P acting through the center of gravity together with a couple
M about the center of gravity. Consider a symmetric shell with
axial moment of inertia A as well as moment of inertia B about
any axis perpendicular to axis of symmetry, through the center of
gravity.
We shall assume that the resultant force P lies in the plane of
incidence and that the axis of the couple M is normal to this plane.
This is not strictly true but is a reasonable approximation for small
yaw. Let α be the angle of incidence, that is, angle between
direction of axis of symmetry and direction of motion (GZ). Let GH
be perpendicular to GZ and resolve P along GH (lift L) and ZG
(drag D), as shown in the following (Fig. 12.1):

184
(Fig. 12.1): Forces acting on a spinning shell

We assume that the couple M is proportional to α and then


for small α, we can write:

M = Mα . α
Where:
Mα = M / α

Note that Mα depends on:


1. Shape of shell
2. Velocity
3. Change in air density

We shall however ignore


1 and 3 for a given shell and treat M as constant.
In order to use the
previous work on tops we need only note one important result
from the theory of dynamics of a rigid body: namely that motion of
and motion about the center of gravity of a rigid body are
independent, that is, can be treated separately.

12. 2 STABILITY OF A PROJECTILE:

We can use the top equations for motion about the center of
gravity, where:

F = Mα / B

We note that the couple Mα. α takes the place of the weight
of the top. Also, certain forces have been neglected, including the
Magnus force, Coriolis force (due to the rotation of the earth),
the spin reducing couple (due to skin friction), and the damping

185
couple (due to the moment of the air resistance).

12.2.1 The Stability Coefficient of a Shell:


The stability coefficient is the same as for the top, that is:
s = A2 N2 / (4 B2 F)
= A2 N2 / (4 B Mα)
We require s > 1. However a more satisfactory value is at
least 1.2, where values smaller than this may give rise to
undesirably large yaw angles following a disturbance. A value of s
around 1.5 is usually favoured.

It should be remembered that it is possible to introduce too


much stabilization. Excessive spin rates cause the shell to
respond rather slowly to changes in trajectory direction and the
axis of the shell tends to remain in its initial direction.

To find N for a spinning shell we have, if the rifling is 1 in n


(that is shell turns one complete revolution in a distance of n
calibers), that the shell turns 2π radians in a distance 2nr, where:
r is the radius of the shell, and if it emerges at a speed V the rate
of turn is:

N = 2 π V / (2 n r)
= π V / (n r) (rad / sec)

and so:
s = A2 N2 / (4 B2 F)
= A2 π 2 V2 / (4 n2 r2 B2 F)
= A2 π 2 V2 / (4 n2 r2 B Mα)

To express the stability coefficient in terms of the more


fundamental properties of the shell and gun, we may express the
yawing moment derivative Mα in terms of the yawing moment

186
coefficient derivative CMα , thus:

Mα = ½ ρ V2 Ao d. CMα
where:
d: is the caliber of the shell.
Then:
s = (8 π / ρ n2 d5) (A2/ B) (1 / CMα)
where:
Ao = ¼ π d2.
It is apparent that the worst condition for the gyroscopic
stability is when ρ and CMα are at a maximum.
We know that atmospheric density ρ is at a maximum at
sea (ground) level, so the value of s will be lowest at launch for
the charge corresponding to the Mach number for which CMα is
maximum.
For a typical shell the yawing moment coefficient derivative
CMα against Mach number curve has the general form as shown
in the following (Fig. 12.2):

(Fig. 12.2): Variation of CMα with Mach number M

187
Thus the gyroscopic stability is least during a transonic
launch, that is, at a muzzle velocity at or near the speed of sound,
approximately 340 (m/s).

As the diameter and mass of the projectile are usually


fixed, as far as the aerodynamics is concerned, A, the axial inertia
is also nearly fixed, hence gyroscopic stability must be attained
mainly by varying (N / V), that is, (2π / n d), B and CMα. Overall,
these reduce to varying the length to diameter ratio of the shell.
This is the reason why there is a maximum length of shell, in
terms of calibers, that can be gyroscopically stabilized with a
given rifling twist.

12.2.2 Angular motion of Spinning Projectiles


(Two-arm Model):
If we refer to our previous analysis of the spinning top we
saw that by looking down on the top the motion could be
interpreted as two arms: one slow and the other fast. This same
procedure can be applied to our spinning projectile and now the
locus of the nose of the spinning projectile can be interpreted in
terms of two rotating arms, OA, AB, hinged at A, O being fixed.

(Fig. 12.3): Two arm model applied to spinning projectile


The slow arm, OA rotates with angular velocity:
AN/ 2B – [(AN/ 2B)2 – (Mα/ B) ]1/2

The fast m, AB rotates with angular velocity:


AN/ B + [(AN/ 2B)2 – (Mα/ B) ]1/2

The rates of rotation may be written more concisely as:


AN/ 2B – n
AN/ 2B + n

The directions shown in (Fig. 12.3) would represent the


angular motion viewed from the rear, the spin being in the
clockwise direction (from the rear).

12.2.3 Some Typical Trochoidal Motions:

The application of the Two-arm model provides, the means


to generate typical angular motions.

The general shape of the locus of the nose is sometimes


termed a trochoid. The particular pattern is determined by:
1. The ratio of (AN/ 2B) to n, that is, the balance between
spin rate and aerodynamic moment
2. The ratio of the two arms, which depends only on the
initial disturbance.

For convenience, we shall denote the lengths of the fast


and slow arms by R, S, respectively.

Typical patterns for an aerodynamically stable projectile


are shown in (Fig. 12.4), where the number of loops in the
complete circle depends on the ratio of (AN/ 2B) to n, and
increases as their ratio approaches one, that is, as the
aerodynamic influence decreases.
(Fig. 12.4): Typical patterns for
aerodynamically stable projectiles

Typical patterns for an aerodynamically unstable projectile


are shown in the following (Fig. 12.5):

(Fig. 12.5): Typical patterns for


aerodynamically unstable projectiles

Again, the number of loops will increases as the ratio of


(AN/ 2B) to n approaches one; that is, as the aerodynamic
influence decreases.

It will be noticed that the looping is outward for the


stable and inward for the unstable projectile, but in both
cases the motion is confined between two circular
boundaries.
12.2.4 Rosette Motion:
The type of motion most typical of a projectile from a gun is
given by: R = S and corresponds to the ellipse degenerating into a
straight line. Such a motion is sometimes called Rosette motion.
The angle of yaw is initially zero and repeatedly passes
through the zero value, as shown in (Fig. 12.6):

(Fig. 12.6): Rosette motion

12.2.5 Steady Precessions:

Appropriate initial conditions could result in a steady


precessional motion, the locus in this case being a circle. Two
precessional motions are possible. If the length of the slow arm is
zero (R = 0) we get the fast precession in a clockwise direction
(for clockwise spin) at rate:
AN/ 2B + n
If the length of the fast arm is zero (S = 0), we get the
slow precession, at rate:

AN/ 2B - n
The direction will be anti-clockwise for the aerodynamically
stable projectile and clockwise for the unstable one.
12.2.6 Equilibrium Yaw Motion:
As discussed here in before, the aerodynamically unstable
projectile will fly with an average angle of yaw to the right of the
trajectory. The trochoidal motion described above is retained,
but its center is displaced by the amount of the equilibrium yaw.
For an aerodynamically stable missile, the equilibrium yaw
will be to the left of the trajectory since the sign of M is reversed.
Typical motions for the two cases are illustrated in (Fig. 12.7):

(Fig. 12.7): Equilibrium yaw motion

12.3 EFFECT OF AERODYNAMIC DAMPING:


In representing the
angular motion in terms of a looping movement between two fixed
circular boundaries, the effect of aerodynamic damping has been
ignored. If damping terms are included in the equations, the
consequences are as follows:
1. The spin rate declines very slowly
2. The fast arm is strongly damped
3. The slow arm is weakly damped for the aerodynamically
stable projectile and weakly negatively damped, that is,
slowly increases, for the unstable projectile.

12.4 EQUILIBRIUM YAW AND DRIFT:


If the trajectory were a
straight line, the spin stabilization would prevent the yaw angle
growing following any disturbances, and damping moments would
in time reduce the amplitude of the motion.
Near the muzzle of the gun the curvature of the trajectory is
small but, as the shell proceeds, the curvature increases, reaching
a peak at the vertex. The spin stabilization then has the effect of
ensuring that the average direction of the shell axis turns with the
trajectory: the axis will of course oscillate about this mean
direction. For the average direction of the shell axis to follow the
curve of the trajectory, the angular momentum vector “ I ” must
slowly rotate. This implies that increments ω of angular
momentum must be continually added to “ I ” in the downward
direction.
Viewed from the top the continuous addition of angular
momentum ω requires a continuous yawing moment M in the
direction shown in (Fig. 12.8):

(Fig. 12.8): Increments of angular momentum


This can only be pro-
duced if the shell flies on the average with a small angle of yaw to
the right of the trajectory. This is called equilibrium yaw.
The increment ω in time
dt may be written as:
ω = I d
where:
: is the inclination of the trajectory to the horizontal.
it follows that:
M = dω/ dt
= I d/ dt
then writing:
M = Mα α
I =AN
the equilibrium yaw a is given by:
α = A N  / Mα
In terms of the stability coefficient s of the shell, the equilibrium
yaw may be expressed as:
α = 4 s B g cos  / (ANV)
We have seen that by
spinning a shell at the appropriate rate, gyroscopic effects can be
used to keep the nose pointing approximately along the trajectory,
as shown in (Fig. 12.9):
(Fig. 12.9): Equilibrium yaw angle

However, the shell inevitably suffers some yaw and the


presence of equilibrium yaw will mean that the nose is pointing to
the right of the trajectory. The air meeting it during flight is thus
deflected to the left, so creating a cross wind force to the right,
which, will mean that the shell will drift to the right in flight. The
drift increases with the time of flight and, for a given muzzle
velocity, the amount is found to vary with the angle of projection
θ, according to the law:

D = C tan θ
Where:
D: is the amount of drift at certain time of flight, for a given
muzzle velocity.
C: is a constant can be found experimentally.

We have now looked at the ballistics involved in launching a


projectile and in its flight to the target. The next stage is
concerned with its performance when it reaches the target. This is
called terminal ballistics and is the subject of another course.
CHAPTER 12
RANDOM VARIABLES
AND PROBABILITY DISTRIBUTIONS

13.1 RANDOM VARIABLES:


Sometimes, we are not interested in the outcomes of a
random experiment, but we are interested in real numbers
connected with these outcomes. For example, we may be
interested in the number of heads when a coin is tossed twice, the
sample space will be:
S= {(H, H), (H, T), (T, H), (T, T)}
then, we define a function:
X: S → R (where: R is the set of real numbers), as follows:
(H, H) → 2
(H, T) → 1
(T, H) → 1
(T, T) → 0
From which we can see that the function X takes the values:
0, 1, and 2 i.e. the range of this function is the set {0, 1, 2}, these
values represent the number of heads when a coin is tossed
twice. This function is called a Random Variable.

Definition of Random Variable:


If S is a sample space of random experiment and R is the set
of real numbers then any function X: S → R is called a random
variable defined on the sample space S, in which the random
variable makes partition for it.

Definition of Discrete Random Variable:


The Random Variable is discrete if its range is a finite or
countable set of real numbers. For Example; the random variable
X which is defined before above is Discrete Random Variable.
Definition of Continuous Random Variable:
The Random Variable is continuous if its range is
uncountable set of real numbers, i.e. open or closed interval. For
example if a point (x, y) is selected at random inside or on the
circle: x2+ y2= 4; and S denote the sample space of this
experiment:
S = {(x, y): x2+ y2 ≤ 4}
Hence; the range of the random variable X is the interval [0, 2].

13.2 PROBABILITY DISTRIBUTIONS:


There are two types of probability distributions according to
the type of random variable itself.

1) Discrete Probability Distributions:


If X is a discrete random variable and its range is:
{x1, x1, x1, x1, ….., x1, ……, x1}
thus the function P(xi) for every i = 1, 2, 3, ….., n; determine the
probability distribution of the discrete random variable X. This
probability distribution can be represented by the set of ordered
pairs {(x1, p(x1)), (x2, p(x2)), (x3, p(x3)), ….., (xn, p(xn))}, and the
function P satisfies the conditions:
a) P(xi) ≥ 0 for every I = 1, 2, 3, ….., n
b) P(x1) + P(x2) + P(x3) + …..+ P(xn) = 1
For example; if a coin is tossed three times in succession,
and the random variable X denote the number of heads occurring;
then the sample space of this experiment consists of; 23= 8
elements i.e. :
S = {(H, H, H), (H, H, T), (H, T, H), (T, H, H), (H, T, T),
(T, H, T), (T, T, H), (T, T, T)}

So the rang of X is the set {0, 1, 2, 3}


13.4 THE MEAN, THE VARIANCE AND,
THE STANDARD DEVIATION OF RANDOM VARIABLE:

The Mean (The Expectation), The Variance and The


Standard Deviation of The Random Variable are summarizing
important characteristics of probability distributions and properties
of populations.
Also these concepts are used to compare between different
populations having the same means with the same units of
measure.
Some distributions do not have a means, some do not have
a variance as well. The mean exists whenever the variance exist,
but not vice versa.

13.4.1 The Mean (Expected Value or Expectation); μ:

The mean of random variable X; gives a single value about


which the observed values of the random variable concentrate.

If a random variable X takes on the values: x1, x2, x3,….., xn


with probabilities: p(x1), p(x2), p(x3), ….. , p(xn) respectively, then:

μ = x1. p(x1) + x2. p(x2) + x3. p(x3) + ….. + xn. p(xn)


= Σ xi. p(xi) (from: i= 1 to: i= n)
If the probabilities of each are equal, i.e: p(xi) = 1/n, then:
μ = (1/n) Σ xi (from: i= 1 to: i= n)
13.4.2 The Variance; {Var(x), σx2, or σ2}:
The Variance of random variable is a measure of
dispersion (scatter) of the values of the random variable about the
mean. In plain language it can be expressed as; the average of
the square of the distance of each data point from the mean i.e. it
is the mean squared deviation, so the variance of a real-valued
random variable is its second central moment:
σ2 = Σ ( xi – μ)2. p(xi) or (from: i= 1 to: i= n)
= [Σ ( xi2. p(xi)] – μ2 (from: i= 1 to: i= n)
If the probabilities of each are equal, i.e: p(xi) = 1/n, then:
σ2 = Σ ( xi – μ)2. (1/n) or (from: i= 1 to: i= n)
= [(1/n).Σ ( xi2)] – μ2 (from: i= 1 to: i= n)

13.4.3 The Standard Deviation; σ:


The Standard Deviation of random variable is also a
measure of dispersion, where it is the measure of the spread of
the values of the random variable:
σ = √ σ2

13.4.4 The Coefficient of Variation:


It is suitable to compare between the dispersion of two
groups having different means of different units of measurement
Coefficient of Variation = (σ/μ) x 100%
13.4.5 The Normal Random Distribution:
A continuous random variable "X" is said to be a Normal
Random Variable if its range is: ]-∞, +∞[ ; and its probability
density function has a bell-shaped curve. It is often referred to as
Gaussian Distribution in honor of Karl Gauss who derived its
equation from a study of errors in repeated measurements of the
same quantity;

# The Probability Density Function Curve

13.4.6 The Standard Normal Random Distribution:


Fortunately, all normal distribution can be reduced to a
standard normal distribution with; μ= 0 and σ= 1.
The total area under the normal curve and above the
horizontal axis is equal to One, and x= μ divides this area into two
equal parts each is equal to 1/2.
So, if "X" is a normal random variable with mean; μ and
variance; σ2 ,and we need to calculate the probability that it lies in
the interval [a, b], we reduce it to the standard normal variable; Z
in order to we can use the standard tables:
Z = (X- μ)/ σ

P ( a ≤ X ≤ b) = P ((a- μ)/ σ ≤ (X- μ)/ σ ≤ (b- μ)/ σ)


In Case of Ammunition Tests:
The probable error must be within the value:
P.E. = 0.6745 σ
If it exist:
(μ - xi) > P.E. x (factor of the number of rounds)
"From tables"
then this round " i " must be excluded from results; and recalculate
the values of; μ and σ.

You might also like