You are on page 1of 212

Module 2

Licence Category
B1 and B2

Physics

2.2 Mechanics

For Training Purposes Only


Intentionally Blank

2-2 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Copyright Notice
© Copyright. All worldwide rights reserved. No part of this publication may be reproduced,
stored in a retrieval system or transmitted in any form by any other means whatsoever: i.e.
photocopy, electronic, mechanical recording or otherwise without the prior written permission of
ST Aerospace Ltd.

Knowledge Levels — Category A, B1, B2, B3 and C Aircraft


Maintenance Licence
Basic knowledge for categories A, B1, B2 and B3 are indicated by the allocation of knowledge levels indicators (1,
2 or 3) against each applicable subject. Category C applicants must meet either the category B1 or the category B2
basic knowledge levels.
The knowledge level indicators are defined as follows:

LEVEL 1
• A familiarisation with the principal elements of the subject.
Objectives:
• The applicant should be familiar with the basic elements of the subject.
• The applicant should be able to give a simple description of the whole subject, using common words and
examples.
• The applicant should be able to use typical terms.

LEVEL 2
• A general knowledge of the theoretical and practical aspects of the subject.
• An ability to apply that knowledge.
Objectives:
• The applicant should be able to understand the theoretical fundamentals of the subject.
• The applicant should be able to give a general description of the subject using, as appropriate, typical
examples.
• The applicant should be able to use mathematical formulae in conjunction with physical laws describing the
subject.
• The applicant should be able to read and understand sketches, drawings and schematics describing the
subject.
• The applicant should be able to apply his knowledge in a practical manner using detailed procedures.

LEVEL 3
• A detailed knowledge of the theoretical and practical aspects of the subject.
• A capacity to combine and apply the separate elements of knowledge in a logical and comprehensive
manner.
Objectives:
• The applicant should know the theory of the subject and interrelationships with other subjects.
• The applicant should be able to give a detailed description of the subject using theoretical fundamentals
and specific examples.
• The applicant should understand and be able to use mathematical formulae related to the subject.
• The applicant should be able to read, understand and prepare sketches, simple drawings and schematics
describing the subject.
• The applicant should be able to apply his knowledge in a practical manner using manufacturer's
instructions.
• The applicant should be able to interpret results from various sources and measurements and apply
corrective action where appropriate.

Module 2.2 Mechanics 2-3


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-4 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Table of Contents

Module 2.2 Mechanics _______________________________________________________ 9


Statics___________________________________________________________________ 9
Mass, Force and Weight ___________________________________________________ 9
Stress, Strain and Hooke’s Law _____________________________________________ 17
Materials Behaviour ______________________________________________________ 23
Nature and Properties of Solids, Liquids and Gas _______________________________ 29
Pressure and Force ______________________________________________________ 35
Barometers ____________________________________________________________ 57
Buoyancy ______________________________________________________________ 59
Buoyancy ______________________________________________________________ 59
Kinetics ________________________________________________________________ 69
Linear Motion ___________________________________________________________ 69
Rotational Motion ________________________________________________________ 79
Periodic Motion _________________________________________________________ 91
Simple Harmonic Motion (SHM) ____________________________________________ 92
Simple Machines and the Principle of Work___________________________________ 101
Dynamics ______________________________________________________________ 115
Newton’s Laws _________________________________________________________ 115
Motion in a Circle _______________________________________________________ 125
Friction _______________________________________________________________ 133
Work, Energy and Power _________________________________________________ 141
Momentum ____________________________________________________________ 153
Torque _______________________________________________________________ 165
The Gyroscope ________________________________________________________ 173
Fluid Dynamics _________________________________________________________ 177
The Atmosphere _______________________________________________________ 177
Density and Specific Gravity ______________________________________________ 187
Compressibility in Fluids _________________________________________________ 195
Viscosity______________________________________________________________ 197
Drag and Streamlining ___________________________________________________ 199
Bernoulli’s Principle _____________________________________________________ 203

Module 2.2 Mechanics 2-5


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-6 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Module 2.2 Enabling Objectives and Certification Statement
Certification Statement
These Study Notes comply with the syllabus of Singapore Airworthiness Requirements Part 66 -
Aircraft Maintenance Licensing:
SAR-66 Licence Category
Objective
Reference B1 B2
Mechanics 2.2

Statics 2.2.1 2 1
Forces, moments and couples, representation
as vectors
Centre of gravity
Elements of theory of stress, strain and
elasticity: tension, compression, shear and
torsion
Nature and properties of solid, fluid and gas
Pressure and buoyancy in liquids
(barometers)
Kinetics 2.2.2 2 1
Linear movement: uniform motion in a straight
line,motion under constant acceleration
(motion under gravity);
Rotational movement: uniform circular motion
(centrifugal/centripetal forces);
Periodic motion: pendular movement;
Simple theory of vibration, harmonics and
resonance;
Velocity ratio, mechanical advantage and
efficiency
Dynamics 2.2.3
(a) 2 1
Mass
Force, inertia, work, power, energy (potential,
kinetic and total energy), heat, efficiency
(b) 2 2
Momentum, conservation of momentum
Impulse
Gyroscopic principles
Friction: nature and effects, coefficient of
friction (rolling resistance)
Fluid dynamics 2.2.4
(a) 2 2
Specific gravity and density
(b) 2 1
Viscosity, fluid resistance, effects of

Module 2.2 Mechanics 2-7


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
streamlining;
Effects of compressibility on fluids
Static, dynamic and total pressure:
Bernoulli’s Theorem, venturi

2-8 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Module 2.2 Mechanics
Statics
Mass, Force and Weight

Mass
In physics. the term for what we have up to now referred to as the amount of substance or
matter is “mass”. A natural unit for mass is the mass of a proton or neutron. This unit has a
special name, the “atomic mass unit” (amu). This unit is useful in those sciences which deal with
atomic and nuclear matter.

In measuring the mass of objects which we encounter daily, this unit is much too small and
therefore very inconvenient. For example, the mass of a bowling ball expressed in amu’s would
be about 4,390,000,000,000,000,000,000,000,000.

One kilogram equals 602,000,000,000,000,000,000,000,000 amu. Since one amu is the mass
of a proton or neutron we know immediately that a kilogram of anything has this combined
number of protons and neutrons contained in it.

The kilogram is the SI unit of mass. In the English system, the standard unit of mass is the slug.

The conversion is:

1 slug = 14.59 kg = 8,789,000,000,000,000,000,000,000,000 amu

We will use the conveniently sized units, the slug in the English system and the kilogram in the
metric system, for all of the problems that we will do in this course. Note that the above
conversion, 1 slug = 14.59 kilogram, is listed with your conversion factors in the table of
conversion factors (Table 1-1).

Force
The physicist uses the word “force” to describe any push or pull. A force is one kind of vector.
A vector is a quantity that has both size and direction.

A force has a certain magnitude or size. Also, a force is always in a certain direction. To
completely describe a force, it is necessary to specify both the size of the push or pull and its
direction.

The units in which force are measured are the pound (lb.) in the English system and the Newton
(N) in the metric system. The Newton is named for Sir Isaac Newton, a famous British physicist
who lived in the 17th century.

The relationship between the metric and English units is given by the conversion factor:

1 lb. = 4.448 N

Module 2.2 Mechanics 2-9


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Weight
A weight is one kind of force. It is defined as the gravitational pull of the earth on a given body.
The direction of this force is toward the geometrical centre of the earth.

Distinction between Mass and Weight


The physicist very carefully distinguishes between “mass” and “weight”. As we have seen,
mass is the quantity of matter, determined by the number of protons and neutrons in the body,
and weight is a measure of the gravitational pull of the earth on this quantity of matter.

It may seem that this is an unimportant distinction. However, there is one important difference.

The mass of an object is the same wherever this object is in the universe. The mass of a stone
is the same if the stone is on the earth, on Mars, in a space ship, or some place in the Milky
Way Galaxy. If the stone is not on the earth but is in a space station orbiting the earth some
distance from the earth’s surface, the weight of this stone is different from its weight on the
earth’s surface. If the stone is on the planet Mars, we speak of its “weight on Mars”, the
gravitational pull of Mars on the stone.

As you have probably figured out, the greater the mass of an object on the surface of the earth,
the greater is the weight of this object. These two quantities are approximately proportional to
each other as long as the body remains on the surface of the earth. The word “approximately” in
the previous sentence refers to the fact that the pull of the earth on a body of a given mass
varies slightly with the position of the body on the earth’s surface. For example, a body that
weighs 57.3 lbs. at the North Pole would weigh 57.0 lbs. at a place on the equator. This occurs
because a body at either pole is slightly closer to the centre of the earth than it is at the equator.
Thus, the pull of the earth on the body is greater at the poles and slightly smaller at other places
on the earth. However, we usually neglect this slight difference.

Physicists and engineers measure masses of bodies in slugs or kilograms and weights in
pounds or Newtons. The equation relating mass and weight is:

w = mg

In this equation, g has a definite numerical value. We will use the following relations:

lbs N
g = 32 or g = 9.8
slug kg

There is a great source of confusion in British marketing practices. For example, we often see
on a packet of sugar the information regarding the contents:

1 kg or 2.2 lbs

We note that 2.2 lbs. equals 1 kg. We have just learned that 2.2 lbs. is the weight of the sugar
and that 1 kg is the mass of the sugar. In other words, British packaging practices list the weight

2-10 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
of the product if we deal with the English system and list the mass of the product if we are in the
metric system.
For example, suppose the weight of a piece of cheese is marked 32 oz. and we wish to know
the number of grams. First we convert the weight in ounces to 2 lbs. Then we convert from
pounds to Newtons.

4.448 N
W = 2 lbs × = 8.90 N
1lb

Next, we use the relation:

W
w = mg or m=
g

Therefore, we write:

8.90 N
m= = 0.908 kg = 908 grams
9.8 N/kg

Note that we can convert from pounds to Newtons since both are units of weight and we can
convert from kilograms to slugs since both are units of mass. However, if we want to find a
mass if we know a weight or a weight if we know a mass we must use the equation:

m = w/g or w = mg

In summary, let us note that mass is a measure of the quantity of matter - ultimately, a measure
of the number of protons and neutrons in the body and weight is the force with which the earth
pulls on a body. These are related but not identical concepts. The units of mass are slugs and
kilograms. The units of weight are pounds and Newtons. A mass can be changed from slugs to
kilograms and vice versa. A weight can be changed from Newtons to pounds or vice versa.
However, one cannot say that one pound equals 454 grams. The only correct statement is that
a body having a weight of one pound has a mass of 454 grams. The equation relating mass and
weight is:

W
w = mg or m=
g

Module 2.2 Mechanics 2-11


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-12 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Problems

1. What is the mass of a body having a weight of 45 N?

2. What is the weight of a body having a mass of 23 kg?

3. What is the mass of a body having a weight of 350 lbs.?

4. What is the weight of a body having a mass of 23.6 slugs?

5. What is the weight (in lbs.) of the corn flakes in a box where the mass is listed as
680 g?

6. What is the mass in grams of 2.5 lbs. of bologna?

Module 2.2 Mechanics 2-13


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-14 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Answers

All answers are to 3 significant figures


1. 4.59 kg
2. 225 N
3. 10.9 slugs
4. 755 lbs.
5. 1.45 lbs.
6. 1140g

Module 2.2 Mechanics 2-15


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-16 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Stress, Strain and Hooke’s Law

Introduction
Structural integrity is a major factor in aircraft design and construction. No production aeroplane
leaves the ground before undergoing extensive analysis of how it will fly, the stresses it will
tolerate and its maximum safe capability.

Every aircraft is subject to structural stress. Stress acts on an aeroplane whether on the ground
or in flight. Stress is defined as a load applied to a unit area of material. Stress produces a
deflection or deformation in the material called strain. Stress is always accompanied by strain.

Current production general aviation aircraft are constructed of various materials, the primary
being aluminium alloys. Rivets, bolts, screws and special bonding adhesives are used to hold
the sheet metal in place. Regardless of the method of attachment of the material, every part of
the fuselage must carry a load, or resist a stress placed on it. Design of interior supporting and
forming pieces, and the outside metal skin all have a role to play in assuring an overall safe
structure capable of withstanding expected loads and stresses.

The stress a particular part must withstand is carefully calculated by engineers. Also, the
material a part is made from is extremely important and is selected by designers based on its
known properties. Aluminium alloy is the primary material for the exterior skin on modern
aircraft. This material possesses a good strength to weight ratio, is easy to form, resists
corrosion, and is relatively inexpensive.

Types of Structural Stress


The five basic structural stresses to which aircraft are subject are:

1. Tension
2. Compression
3. Torsion
4. Shear
5. Bending

While there are many other ways to describe the actual stresses which an aircraft undergoes in
normal (or abnormal) operation, they are special arrangements of these basic ones.

Tension - is the stress acting against another force that is trying to pull something apart. For
example, while in straight and level flight the engine power and propeller are pulling the
aeroplane forward. The wings, tail section and fuselage, however, resist that movement
because of the airflow around them. The result is a stretching effect on the airframe. Bracing
wires in an aircraft are usually in tension.

Compression - is a squeezing or crushing force that tries to make parts smaller. Anti-
compression design resists an inward or crushing force applied to a piece or assembly. Aircraft
wings are subjected to compression stresses. The ability of a material to meet compression
requirements is measured in pounds per square inch (PSI).

Module 2.2 Mechanics 2-17


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Torsion - is a twisting force. Because aluminium is used almost exclusively for the outside, and,
to a large extent, inside fabrication of parts and covering, its tensile strength (capability of being
stretched) under torsion is very important. Tensile strength refers to the measure of strength in
pounds per square inch (PSI) of the metal. Torque (also a twisting force) works against torsion.
The torsional strength of a material is its ability to resist torque. While in flight, the engine power
and propeller twist the forward fuselage. The force, however, is resisted by the assemblies of
the fuselage. The airframe is subjected to variable torsional stresses during turns and other
manoeuvres.

Shear - stress tends to slide one piece of material over another. Consider the aircraft fuselage.
The aluminium skin panels are riveted to one another. Shear forces try to make the rivets fail
under flight loads; therefore, selection of rivets with adequate shear resistance is critical. Bolts
and other fasteners are often loaded in shear, an example being bolts that fasten the wing to
the spar or carry-through structure. Although other forces may also be present, shear forces try
to rip the bolt in two. Generally, shear strength is less than tensile or compressive strength in a
particular material.

Bending - is a combination of two forces, compression and tension. During bending stress, the
material on the inside of the bend is compressed and the outside material is stretched in
tension. An example of this is the G-loading an aeroplane structure experiences during
manoeuvring. During an abrupt pull-up, the aeroplane's wing spars, wing skin and fuselage
undergo positive loading and the upper surfaces are subject to compression, while the lower
wing skin experiences tension loads. There are many other areas of the airframe structure that
experience bending forces during normal flight.

An aircraft structure in flight is subjected to many and varying stresses due to the varying loads
that may be imposed. The designer's problem is trying to anticipate the possible stresses that
the structure will have to endure, and to build it sufficiently strong to withstand these. The
problem is complicated by the fact that an aeroplane structure must be light as well as strong.

Stress, Strain and Young’s Modulus


What is known as Axial (or Normal) Stress, is defined as the force perpendicular to the cross
sectional area of the member divided by the cross sectional area. Or

Stress =
Force
Area
(
units lb/in2 or N/m2 )
In figure 2-1, a solid rod of length L, is under simple tension due to force F, as shown. If we
divide that axial force, F, by the cross sectional area of the rod (A), this would be the axial stress
in the member. Axial stress is the equivalent of pressure in a gas or liquid. As you remember,
pressure is the force/unit area. So axial stress is really the 'pressure' in a solid member. Now
the question becomes, how much 'pressure' can a material bear before it fails.

2-18 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Figure 2.1: Tensile Stress

In fact, if we look at a metal rod in simple tension as shown in figure 2.1, we see that there will
be an elongation (or deformation) due to the tension. If we then graph the tension (force) verses
the deformation we obtain a result as shown in figure 2.2.

Figure 2.2: Force-Extension diagram

In figure 2.2, we see that, if our metal rod is tested by increasing the tension in the rod, the
deformation increases. In the first region the deformation increases in proportion to the force.
That is, if the amount of force is doubled, the amount of deformation is doubled. This is a form
of Hooke's Law and could be written this way: F ∞ k (deformation), where k is a constant
depending on the material (and is sometimes called the spring constant). After enough force
has been applied the material enters the plastic region - where the force and the deformation
are not proportional, but rather a small amount of increase in force produces a large amount of
deformation. In this region, the rod often begins to 'neck down', that is, the diameter becomes
smaller as the rod is about to fail. Finally the rod actually breaks.

The point at which the Elastic Region ends is called the elastic limit, or the proportional limit. In

Module 2.2 Mechanics 2-19


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
actuality, these two points are not quite the same. The Elastic Limit is the point at which
permanent deformation occurs, that is, after the elastic limit, if the force is taken off the sample,
it will not return to its original size and shape, permanent deformation has occurred. The
Proportional Limit is the point at which the deformation is no longer directly proportional to the
applied force (Hooke's Law no longer holds). Although these two points are slightly different, we
will treat them as the same in this course.

Next, rather than examining the applied force and resulting deformation, we will instead graph
the axial stress verses the axial strain (figure 2-2). We have defined the axial stress earlier.

The axial strain is defined as the fractional change in length or

Strain = (deformation of member) divided by the (original length of member)

∆L
Strain =
Lo

Figure 2.3: Axial force in a member of length


Lo causing deformation (extension) of ∆L
We may write:

Deformatio n
Strain =
Original Length

where Lo is the original length of the member.

Strain has no units - since its length divided by length, however it is sometimes expressed as
'in/in (or inches per inch)' in some texts.

As we see from figure 2.4, the Stress verses Strain graph has the same shape and regions as
the force verses deformation graph in figure 2.2. In the elastic (linear) region, since stress is
directly proportional to strain, the ratio of stress/strain will be a constant (and actually equal to
the slope of the linear portion of the graph).

This constant is known as Young's Modulus, and is usually symbolized by an E or Y. We will


use E for Young's modulus. We may now write

2-20 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Stress
Young's Modulus (E) =
Strain

(This is another form of Hooke's Law.)

Figure 2.4: Stress-Strain graph

The value of Young's modulus - which is a measure of the amount of force needed to produce a
unit deformation - depends on the material.

Young's Modulus for Steel is 30 x 106 lb/in2, for Aluminium E = 10 x 106 lb/in2, and for Brass
E = 15 x 106 lb/in2.

To summarize our stress/strain/Hooke's Law relationships up to this point, we have:

Stress =
Force
Area
(
units lb/in2 or N/m2 )

Deformatio n
Strain =
Original Length

Stress
Young's Modulus =
Strain

Module 2.2 Mechanics 2-21


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Related Definitions

Bulk Modulus - The bulk modulus gives the change in volume of a solid substance as the
pressure on it is changed. The formula for bulk modulus is very similar to that for Young’s
Modulus:
F
Bulk Modulus (B) =
Pressure
= A = Pressure × Vo
Volumetric Strain ∆V ∆V
Vo

Some examples of Bulk Modulus for different materials are given on the next page.

Poisson’s Ratio - As a member is stressed in tension, its length increases (axial strain) and its
width decreases (transverse strain). Poisson’s Ratio is the ratio of transverse strain to the axial
strain in a stressed member.

Cantilever - Figure 2.5 illustrates a cantilever structure. The beam is under bending stress
(which is greatest at the root end) and shear stress (which is constant along the beam).

Figure 2.5: A cantilever structure

2-22 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Materials Behaviour

Elastic - Material deforms under stress but returns to its original size and shape when the
stress is released. There is no permanent deformation. Some elastic strain, like in a rubber
band, can be large, but in metals it is usually small.

Brittle - Material deforms by fracturing. Glass is typically brittle.

Ductile - Material deforms without breaking. Metals and most plastics are ductile.

Viscous - Materials that deform steadily under stress. Purely viscous materials like liquids
deform under even the smallest stress. Even metals may behave like viscous materials under
high temperature and pressure. This is known as creep and affects plastics far more than
metals.

Module 2.2 Mechanics 2-23


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
E

ULTIMATE
STRESS

Table 2.2: Elastic Limit and Ultimate Stress


of some common materials

Table 2.1: Young’s Modulus and Bulk


Modulus of some common materials

EXAMPLE:

The elastic limit for copper is 2.3 x 104 lb/in2 and the ultimate strength is
4.9 x 104 lb/in2 . Suppose that a copper rod has a cross-sectional area of 0.5 in2. A force
of 11,500 lbs. applied longitudinally to this rod would just be within the elastic limit. A force
of 12,000 lbs. would deform the rod in such a way that it would not return to it original size
after the force is removed. A force of 24,500 lbs. would cause the rod to rupture.

2-24 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Problems

1. A steel bolt with a cross-sectional area of 0.1 in2 and a length of 6.0” is subjected to a
force of 580 lbs. What is the increase in length of the bolt?
(Hint: Find the stress. Then use Young’s Modulus of steel to find the strain. From the
strain find the extension)

2. An iron body of volume 145 in3 is subjected to a pressure of 500 lb/in2 . What is the
decrease in volume of this body?

3. A copper rod has a cross-sectional area of 0.04 in2 and a length of 24”. What longitudinal
force must be applied to cause this rod to stretch by 0.0024 in?

4. An aluminium brace inside a wing of a plane has a cross-sectional area of 0.2 in2. What
is the greatest longitudinal force that can be applied to the brace without causing the
brace to be permanently deformed?

Module 2.2 Mechanics 2-25


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-26 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Answers

1. 0.0012 in.
2. 0.05 in3
3. 64 lb.
4. 3800 lb.

Module 2.2 Mechanics 2-27


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-28 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Nature and Properties of Solids, Liquids and Gas
All matter exists in one of three states – solid, liquid or gas. The following notes characterize
the three states:

Solid

• The greatest forces of attraction are between the


particles in a solid and they pack together in a neat and
ordered arrangement.
• The particles are too strongly held together to allow
movement from place to place but the particles vibrate about
there position in the structure.
• With increase in temperature, the particles vibrate faster
and more strongly as they gain kinetic energy.

Figure 2.6: Atom


arrangement in a solid

The properties of a Solid

• Solids have the greatest density (‘heaviest’) because the particles are closest together.
• Solids cannot flow freely like gases or liquids because the particles are strongly held in
fixed positions.
• Solids have a fixed surface and volume (at a particular temperature) because of the
strong particle attraction.
• Solids are extremely difficult to compress because there is no real ‘empty’ space
between the particles.
• Solids will expand a little on heating but nothing like as much as liquids because of the
greater particle attraction restricting the expansion (contract on cooling). The expansion
is caused by the increased strength of particle vibration.

Module 2.2 Mechanics 2-29


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Liquid

• Much greater forces of attraction between the particles in a


liquid compared to gases, but not quite as much as in solids.
• Particles quite close together but still arranged at random
throughout the container, there is a little close range order as you
can get clumps of particles clinging together temporarily.
• Particles moving rapidly in all directions but more frequently
colliding with each other than in gases.
• With increase in temperature, the particles move faster as
they gain kinetic energy.
Figure 2.7: Atom
arrangement in a liquid
Properties of a Liquid

• Liquids have a much greater density than gases (‘heavier’) because the particles are
much closer together.
• Liquids flow freely despite the forces of attraction between the particles but liquids are
not as ‘fluid’ as gases.
• Liquids have a surface, and a fixed volume (at a particular temperature) because of the
increased particle attraction, but the shape is not fixed and is merely that of the container
itself.
• Liquids are not readily compressed because of the lack of ‘empty’ space between the
particles.
• Liquids will expand on heating (contract on cooling) but nothing like as much as gases
because of the greater particle attraction restricting the expansion. When heated, the
liquid particles gain kinetic energy and hit the sides of the container more frequently, and
more significantly, they hit with a greater force, so in a sealed container the pressure
produced can be considerable.

2-30 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Gas

• Almost no forces of attraction between the particles which are


completely free of each other.
• Particles widely spaced and scattered at random throughout the
container so there is no order in the system.
• Particles moving rapidly in all directions, frequently colliding with
each other and the side of the container.
• With increase in temperature, the particles move faster as they
gain kinetic energy.

Figure 2.8: Atom


arrangement in a gas

Properties of a Gas

• Gases have a low density (‘light’) because the particles are so spaced out in the
container (density = mass ÷ volume).
• Gases flow freely because there are no effective forces of attraction between the
particles.
• Gases have no surface, and no fixed shape or volume, and because of lack of particle
attraction, they spread out and fill any container.
• Gases are readily compressed because of the ‘empty’ space between the particles.
• If the ‘container’ volume can change, gases readily expand on heating because of the
lack of particle attraction, and readily contract on cooling. On heating, gas particles gain
kinetic energy and hit the sides of the container more frequently, and more significantly,
they hit with a greater force. Depending on the container situation, either or both of the
pressure or volume will increase (reverse on cooling).
• The natural rapid and random movement of the particles means that gases readily
‘spread’ or diffuse. Diffusion is fastest in gases where there is more space for them to
move and the rate of diffusion increases with increase in temperature.

Module 2.2 Mechanics 2-31


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Changes of State
We can use the diagrams shown below, to explain changes of state and the energy changes
involved.

Figure 2.9: Gas to liquid transformation

Evaporation and Boiling (liquid to gas)


In evaporation and boiling the highest kinetic energy molecules can ‘escape’ from the attractive
forces of the other liquid particles. The particles lose any order and become completely free.
Energy is needed to overcome the attractive forces in the liquid and is taken in from the
surroundings. This means heat is taken in (endothermic). Boiling is rapid evaporation at a fixed
temperature called the boiling point and requires continuous addition of heat. Evaporation
takes place more slowly at any temperature between the melting point and boiling point and
results in the liquid becoming cooler.

Condensing (gas to liquid)


On cooling, gas particles lose kinetic energy and eventually become attracted together to form a
liquid. There is an increase in order as the particles are much closer together and can form
clumps of molecules. The process requires heat to be lost to the surroundings i.e. heat given
out, so condensation is exothermic.

Figure 2.10: Liquid to solid transformation

Melting (solid to liquid)


When a solid is heated the particles vibrate more strongly and the particle attractive forces are
weakened. Eventually, at the melting point, the attractive forces are too weak to hold the
structure together and the solid melts. The particles become free to move around and lose their
ordered arrangement. Energy is needed to overcome the attractive forces, so heat is taken in
from the surroundings and melting is an endothermic process.

2-32 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Freezing (liquid to solid)
On cooling, liquid particles lose kinetic energy and become more strongly attracted to each
other. Eventually at the freezing point the forces of attraction are sufficient to remove any
remaining freedom and the particles come together to form the ordered solid arrangement.
Since heat must be removed to the surroundings freezing is an exothermic process.

Figure 2.11: Phase changes

Summary

Figure 2.12: Solid, liquid and gas summaries

Module 2.2 Mechanics 2-33


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-34 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Pressure and Force

Introduction
The terms force and pressure are used extensively in the study of fluids. It is essential that we
distinguish between the terms. Force means a total push or pull. It is the push or pull exerted
against the total area of a particular surface and is expressed in pounds or grams. Pressure
means the amount of push or pull (force) applied to each unit area of the surface and is
expressed in pounds per square inch (lb/in2) or Newtons per square meter (N/m2). Pressure
maybe exerted in one direction, in several directions, or in all directions.

Computing Force, Pressure, and Area


A formula is used in computing force, pressure, and area in fluid power systems. In this formula,
P refers to pressure, F indicates force, and A represents area. Force equals pressure times
area. Thus, the formula is written

F=PxA

Pressure equals force divided by area. By rearranging the formula, this statement may be
condensed into

F
P=
A

Since area equals force divided by pressure, the formula is written

F
A=
P

Figure 2.13: The Pressure, Force


and Area equation finder

Figure 2.13 illustrates a memory device for recalling the different variations of this formula. Any
letter in the triangle may be expressed as the product or quotient of the other two, depending on
its position within the triangle. For example, to find area, consider the letter A as being set off to
itself, followed by an equal sign. Now look at the other two letters. The letter F is above the
letter P; therefore,

Module 2.2 Mechanics 2-35


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
F
A=
P

NOTE: Sometimes the area may not be expressed in square units. If the surface is rectangular,
you can determine its area by multiplying its length (say, in inches) by its width (also in inches).
The majority of areas you will consider in these calculations are circular in shape. Either the
radius or the diameter may be given, but you must know the radius in inches to find the area.
The radius is one-half the diameter. To determine the area, use the formula for finding the area
of a circle. This is written A = πr2, where A is the area, π is 3.1416 (3.14 or 22/7 for most
calculations), and r2 indicates the radius squared.

Atmospheric Pressure
Recall that the atmosphere is the entire mass of air that surrounds the earth. While it extends
upward for about 500 miles, the section of primary interest is the portion that rests on the earth’s
surface and extends upward for about 7 1/2 miles. This layer is called the troposphere.

If a column of air 1-inch square extending all the way to the "top" of the atmosphere could be
weighed, this column of air would weigh approximately 14.7 pounds at sea level. Thus,
atmospheric pressure at sea level is approximately 14.7 PSI.

As one ascends, the atmospheric pressure decreases by approximately 1.0 PSI for every 2,343
feet. However, below sea level, in excavations and depressions, atmospheric pressure
increases. Pressures under water differ from those under air only because the weight of the
water must be added to the pressure of the air.

Atmospheric pressure can be measured by any of several methods. The common laboratory
method uses the mercury column barometer. The height of the mercury column serves as an
indicator of atmospheric pressure. At sea level and at a temperature of 20° Celsius (C), the
height of the mercury column is 29.92 inches, or 760 millimetres. This represents a pressure of
approximately 14.7 PSI. The 30-inch column is used as a reference standard.

Another device used to measure atmospheric pressure is the aneroid barometer. The aneroid
barometer uses the change in shape of an evacuated metal cell to measure variations in
atmospheric pressure (figure 2.14). The thin metal of the aneroid cell moves in or out with the
variation of pressure on its external surface. This
movement is transmitted through a system of levers to a
pointer, which indicates the pressure.

The atmospheric pressure does not vary uniformly with


altitude. It changes more rapidly at lower altitudes
because of the compressibility of the air, which causes
the air layers close to the earth’s surface to be
compressed by the air masses above them. This effect,
however, is partially counteracted by the contraction of
the upper layers due to cooling. The cooling tends to
increase the density of the air.

Figure 2.14: Aneroid barometer

2-36 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Atmospheric pressures are quite large, but in most instances practically the same pressure is
present on all sides of objects so that no single surface is subjected to a great load.

Atmospheric pressure acting on the surface of a liquid (figure 2.15, view A) is transmitted
equally throughout the liquid to the walls of the container, but is balanced by the same
atmospheric pressure acting on the outer walls of the container. In view B of figure 2.15,
atmospheric pressure acting on the surface of one piston is balanced by the same pressure
acting on the surface of the other piston. The different areas of the two surfaces make no
difference, since for a unit of area, pressures are balanced.

Transmission of Forces Through Liquids


When the end of a solid bar is struck, the main force of the blow is carried straight through the
bar to the other end (figure 2.16, view A). This happens because the bar is rigid. The direction
of the blow almost entirely determines the direction of the transmitted force.

Figure 2.16: Forces acting on solids


and liquids

Figure 2.15: Forces acting on liquids

When a force is applied to the end of a column of confined liquid (figure 2.16, view B), it is
transmitted straight through to the other end and also equally and undiminished in every
direction throughout the column—forward, backward, and sideways—so that the containing
vessel is literally filled with pressure.

Module 2.2 Mechanics 2-37


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
An example of this distribution of force is illustrated in figure 2.17. The outward push of the
water is equal in every direction.

So far we have explained the effects of atmospheric pressure on


liquids and how external forces are distributed through liquids. Let
us now focus our attention on forces generated by the weight of
liquids themselves. To do this, we must first discuss density,
specific gravity, and Pascal’s law.

Figure 2.17: Flat and water


filled water hoses

Pascal’s Law

The foundation of modern hydraulics was established when


Pascal discovered that pressure in a fluid acts equally in all
directions. This pressure acts at right angles to the
containing surfaces. If some type of pressure gauge, with an
exposed face, is placed beneath the surface of a liquid
(figure 2.18) at a specific depth and pointed in different
directions, the pressure will read the same. Thus, we can
say that pressure in a liquid is independent of direction.

Pressure due to the weight of a liquid, at any level, depends


on the depth of the fluid from the surface. If the exposed
face of the pressure gauges, are moved closer to the
surface of the liquid, the indicated pressure will be less.
Figure 2.18: Pascals Law When the depth is doubled, the indicated pressure is
doubled. Thus the pressure in a liquid is directly proportional to the depth.

Consider a container with vertical sides (fig. 2.19) that is 1 foot long and 1 foot wide. Let it be
filled with water 1 foot deep, providing 1 cubic foot of water. We learned earlier in this chapter
that 1 cubic foot of water weighs 62.4 pounds. Using this information and the equation, P = F/A,
we can calculate the pressure on the bottom of the container.

F
P=
A

62.4 lb
=
1 ft 2

=62.4 lb/ft2

2-38 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Since there are 144 square inches in 1 square foot,

62.4
P= = 0.433 lb/in2
144

This can be stated as follows: the weight of a column of water 1 foot high, having a cross-
sectional area of 1 square inch, is 0.433 pound. If the depth of the column is tripled, the weight
of the column will be 3 x 0.433, or 1.299 pounds, and the pressure at the bottom will be 1.299
lb/in2 (PSI), since pressure equals the force divided by the area. Thus, the pressure at any
depth in a liquid is equal to the weight of the column of liquid at that depth divided by the cross-
sectional area of the column at that depth. The volume of a liquid that produces the pressure is
referred to as the fluid head of the liquid. The pressure of a liquid due to its fluid head is also
dependent on the density of the liquid.

If we let A equal any cross-sectional area of a liquid column and h equal the depth of the
column, the volume becomes Ah. Using the equation for density, D = Weight (W) / Volume (V),
the weight of the liquid above area A is equal to AhD.

W W
D= =
V Ah

W = Ahd

Water weight = 62.4 lb.

Water pressure = 62.4 lb/ft2

0.433 lb/in2

Figure 2.19: A body of water

Since pressure is equal to the force per unit area, set A equal to 1. Then the formula pressure
becomes

P =hD

Module 2.2 Mechanics 2-39


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
It is essential that h and D be expressed in similar units. That is, if D is expressed in pounds per
cubic foot, the value of h must be expressed in feet. If the desired pressure is to be expressed
in pounds per square inch, the pressure formula, becomes

hd
P =
144

Pressure and Force in Fluid Power Systems


Pascal was also the first to prove by experiment that the shape and volume of a container in no
way alters pressure. Thus in figure 2.19, if the pressure due to the weight of the liquid at a point
on horizontal line H is 8 PSI, the pressure is 8 PSI everywhere at level H in the system. The
equation P=F/A also shows that the pressure is independent of the shape and volume of a
container. If there is a resistance on the output piston and the input piston is pushed downward,
a pressure is created through the fluid, which acts equally at right angles to surfaces in all parts
of the container. If force 1 is 100 pounds and the area of the input piston is 10 square inches,
then the pressure in the fluid is 10 PSI

Figure 2.20: Hydrostatic pressure in different shaped containers

2-40 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Figure 2.21: Forces, pistons and pressure
Recall that, according to Pascal’s law, any force applied to a confined fluid is transmitted in all
directions throughout the fluid.

NOTE: Fluid pressure cannot be created without resistance to flow. In this case, resistance is
provided by the equipment to which the output piston is attached. The force of resistance acts
against the top of the output piston. The pressure created in the system by the input piston
pushes on the underside of the output piston with a force of 10 pounds on each square inch.

Module 2.2 Mechanics 2-41


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Figure 2.22: Pressure due to forces on pistons

In this case, the fluid column has a uniform cross section, so the area of the output piston is the
same as the area of the input piston, or 10 square inches. Therefore, the upward force on the
output piston is 100 pounds (10 PSI x 10 sq. in.), the same as the force applied to the input
piston. All that was accomplished in this system was to transmit the 100-pound force around the
bend. However, this principle underlies practically all mechanical applications of fluid power.

At this point you should note that since Pascal’s law is independent of the shape of the
container, it is not necessary that the tube connecting the two pistons have the same cross-
sectional area of the pistons. A connection of any size, shape, or length will do, as long as an
unobstructed passage is provided. Therefore, the system shown in figure 2.22, with a relatively
small, bent pipe connecting two cylinders, will act exactly the same as the system shown in
figure 2.21.

2-42 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
The Hydraulic Ram Principle
Consider the situation in figure 2.23, where the input piston is much smaller than the output
piston. Assume that the area of the input piston is 2 square inches. With a resistant force on the
output piston a downward force of 20 pounds acting on the input piston creates a pressure 10
PSI in the fluid.

Figure 2.23: The hydraulic ram principle

Although this force is much smaller than the force applied in figures 2.21 and 2.22, the pressure
is the same. This is because the force is applied to a smaller area.

This pressure of 10 PSI acts on all parts of the fluid container, including the bottom of the output
piston. The upward force on the output piston is 200 pounds (10 pounds of pressure on each
square inch). In this case, the original force has been multiplied tenfold while using the same
pressure in the fluid as before. In any system with these dimensions, the ratio of output force to
input force is always ten to one, regardless of the applied force. For example, if the applied
force of the input piston is 50 pounds, the pressure in the system will be 25 PSI. This will
support a resistant force of 500 pounds on the output piston.

The system works the same in reverse. If we change the applied force and place a 200-pound
force on the output piston (figure 2.23), making it the input piston, the output force on the input
piston will be one-tenth the input force, or 20 pounds. (Sometimes such results are desired.)
Therefore, if two pistons are used in a fluid power system, the force acting on each piston is
directly proportional to its area, and the magnitude of each force is the product of the pressure
and the area of each piston.

Module 2.2 Mechanics 2-43


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Note the white arrows at the bottom of figure 2.23 that indicate up and down movement. The
movement they represent will be explained later in the discussion of volume and distance
factors.

Differential Areas

Figure 2.24: Pressures on unequal areas

Consider the special situation shown in figure 2.24. Here, a single piston (1) in a cylinder (2) has
a piston rod (3) attached to one of its sides. The piston rod extends out of one end of the
cylinder. Fluid under pressure is admitted equally to both ends of the cylinder. The opposed
faces of the piston (1) behave like two pistons acting against each other. The area of one face is
the full cross-sectional area of the cylinder, say 6 square inches, while the area of the other face
is the area of the cylinder minus the area of the piston rod, which is 2 square inches. This
leaves an effective area of 4 square inches on the right face of the piston. The pressure on both
faces is the same, in this case, 20 PSI. Applying the rule just stated, the force pushing the
piston to the right is its area times the pressure, or 120 pounds (20 x 6). Likewise, the force
pushing the piston to the left is its area times the pressure, or 80 pounds (20 x 4). Therefore,
there is a net unbalanced force of 40 pounds acting to the right, and the piston will move in that
direction. The net effect is the same as if the piston and the cylinder had the same cross-
sectional area as the piston rod.

2-44 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Volume and Distance Factors
You have learned that if a force is applied to a system and the cross-sectional areas of the input
and output pistons are equal, as in figures 2.21 and 2.22, the force on the input piston will
support an equal resistant force on the output piston. The pressure of the liquid at this point is
equal to the force applied to the input piston divided by the piston’s area. Let us now look at
what happens when a force greater than the resistance is applied to the input piston.

In the system illustrated in figure 2.23, assume that the resistance force on the output piston is
100 PSI. If a force slightly greater than 100 pounds is applied to the input piston, the pressure in
the system will be slightly greater than 10 PSI. This increase in pressure will overcome the
resistance force on the output piston. Assume that the input piston is forced downward 1 inch.
The movement displaces 10 cubic inches of fluid. The fluid must go somewhere. Since the
system is closed and the fluid is practically incompressible, the fluid will move to the right side of
the system. Because the output piston also has a cross-sectional area of 10 square inches, it
will move 1 inch upward to accommodate the 10 cubic inches of fluid. You may generalize this
by saying that if two pistons in a closed system have equal cross-sectional areas and one piston
is pushed and moved, the other piston will move the same distance, though in the opposite
direction. This is because a decrease in volume in one part of the system is balanced by one
equal increase in volume in another part of the system.

Apply this reasoning to the system in figure 2.23. If the input piston is pushed down a distance
of 1 inch, the volume of fluid in the left cylinder will decrease by 2 cubic inches. At the same
time, the volume in the right cylinder will increase by 2 cubic inches. Since the diameter of the
right cylinder cannot change, the piston must move upward to allow the volume to increase. The
piston will move a distance equal to the volume increase divided by the surface area of the
piston (equal to the surface area of the cylinder). In this example, the piston will move one-tenth
of an inch (2 cu. in. ÷ 20 sq. in.). This leads to the second basic rule for a fluid power system
that contains two pistons: The distances the pistons move are inversely proportional to the
areas of the pistons. Or more simply, if one piston is smaller than the other, the smaller piston
must move a greater distance than the larger piston any time the pistons move.

Module 2.2 Mechanics 2-45


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Relationship between Force, Pressure, and Head
In dealing with fluids, forces are usually considered in relation to the areas over which they are
applied. As previously discussed, a force acting over a unit area is a pressure, and pressure
can alternately be stated in pounds per square inch or in terms of head, which is the vertical
height of the column of fluid whose weight would produce that pressure.

In most of the applications of fluid power, applied forces greatly outweigh all other forces, and
the fluid is entirely confined. Under these circumstances it is customary to think of the forces
involved in terms of pressures. Since the term head is encountered frequently in the study of
fluid power, it is necessary to understand what it means and how it is related to pressure and
force.

At this point you need to review some terms in general use. “Gravity head”, when it is important
enough to be considered, is sometimes referred to as simply “head”. The effect of atmospheric
pressure is referred to as “atmospheric pressure”. (Atmospheric pressure is frequently and
improperly referred to as suction.) Inertia effect, because it is always directly related to velocity,
is usually called “velocity head”; and friction, because it represents a loss of pressure or head, is
usually referred to as “friction head”.

Static and Dynamic Factors


Gravity, applied forces, and atmospheric pressure are static factors that apply equally to fluids
at rest or in motion, while inertia and friction are dynamic factors that apply only to fluids in
motion. The mathematical sum of gravity, applied force, and atmospheric pressure is the static
pressure obtained at any one point in a fluid at any given time. Static pressure exists in addition
to any dynamic factors that may also be present at the same time.

Remember, Pascal’s law states that a pressure set up in a fluid acts equally in all directions and
at right angles to the containing surfaces. This covers the situation only for fluids at rest or
practically at rest. It is true only for the factors making up static head. Obviously, when velocity
becomes a factor it must have a direction, and as previously explained, the force related to the
velocity must also have a direction, so that Pascal’s law alone does not apply to the dynamic
factors of fluid power.

The dynamic factors of inertia and friction are related to the static factors. Velocity head and
friction head are obtained at the expense of static head. However, a portion of the velocity head
can always be reconverted to static head. Force, which can be produced by pressure or head
when dealing with fluids, is necessary to start a body moving if it is at rest, and is present in
some form when the motion of the body is arrested; therefore, whenever a fluid is given velocity,
some part of its original static head is used to impart this velocity, which then exists as velocity
head.

2-46 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Operation of Hydraulic Components
To transmit and control power through pressurized fluids, an arrangement of inter-connected
components is required. Such an arrangement is commonly referred to as a system. The
number and arrangement of the components vary from system to system, depending on the
particular application. In many applications, one main system supplies power to several
subsystems, which are sometimes referred to as circuits. The complete system may be a small
compact unit; more often, however, the components are located at widely separated points for
convenient control and operation of the system.

The basic components of a fluid power system are essentially the same, regardless of whether
the system uses a hydraulic or a pneumatic medium. There are five basic components used in a
system.

These basic components are as follows:

• Reservoir or receiver
• Pump or compressor
• Lines (pipe, tubing, or flexible hose)
• Directional control valve
• Actuating device

Several applications of fluid power require only a simple system; that is, a system which uses
only a few components in addition to the five basic components. A few of these applications are
presented in the following paragraphs. We will explain the operation of these systems briefly at
this time so you will know the purpose of each component and can better understand how
hydraulics is used in the operation of these systems.

Hydraulic Jack
The hydraulic jack is perhaps one of the simplest forms of a fluid power system. By moving the
handle of a small device, an individual can lift a load weighing several tons. A small initial force
exerted on the handle is transmitted by a fluid to a much larger area. To understand this better,
study figure 2.25. The small input piston has an area of 5 square inches and is directly
connected to a large cylinder with an output piston having an area of 250 square inches. The
top of this piston forms a lift platform. If a force of 25 pounds is applied to the input piston, it
produces a pressure of 5 PSI in the fluid, that is, of course, if a sufficient amount of resistant
force is acting against the top of the output piston. Disregarding friction loss, this pressure
acting on the 250 square inch area of the output piston will support a resistance force of 1,250
pounds. In other words, this pressure could overcome a force of slightly under 1,250 pounds.
An input force of 25 pounds has been transformed into a working force of more than half a ton;
however, for this to be true, the distance travelled by the input piston must be 50 times greater
than the distance travelled by the output piston. Thus, for every inch that the input piston
moves, the output piston will move only one-fiftieth of an inch.

Module 2.2 Mechanics 2-47


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
This would be ideal if the output
piston needed to move only a short
distance. However, in most instances,
the output piston would have to be
capable of moving a greater distance
to serve a practical application. The
device shown in figure 2.25 is not
capable of moving the output piston
farther than that shown; therefore,
some other means must be used to
raise the output piston to a greater
height.

Figure 2.25: A simple hydraulic jack

The output piston can be raised


higher and maintained at this
height if additional components
are installed as shown in figure
2.26. In this illustration the jack is
designed so that it can be raised,
lowered, or held at a constant
height. These results are attained
by introducing a number of
valves and also a reserve supply
of fluid to be used in the system.

Notice that this system contains Figure 2.26 (A): A hydraulic jack with valves
the five basic components—the
reservoir; cylinder 1, which
serves as a pump; valve 3, which
serves as a directional control
valve; cylinder 2, which serves as
the actuating device; and lines to
transmit the fluid to and from the
different components. In addition,
this system contains two valves,
1 and 2, whose functions are
explained in the following
discussion.

As the input piston is raised (fig.


2.26, view A), valve 1 is closed
by the back pressure from the Figure 2.26 (B): A hydraulic jack with valves

2-48 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
weight of the output piston. At the same time, valve 2 is opened by the head of the fluid in the
reservoir. This forces fluid into cylinder 1. When the input piston is lowered (fig. 2.26, view B), a
pressure is developed in cylinder 1. When this pressure exceeds the head in the reservoir, it
closes valve 2. When it exceeds the back pressure from the output piston, it opens valve 1,
forcing fluid into the pipeline. The pressure from cylinder 1 is thus transmitted into cylinder 2,
where it acts to raise the output piston with its attached lift platform. When the input piston is
again raised, the pressure in cylinder 1 drops below that in cylinder 2, causing valve 1 to close.
This prevents the return of fluid and holds the output piston with its attached lift platform at its
new level. During this stroke, valve 2 opens again allowing a new supply of fluid into cylinder 1
for the next power (downward) stroke of the input piston. Thus, by repeated strokes of the input
piston, the lift platform can be progressively raised. To lower the lift platform, valve 3 is opened,
and the fluid from cylinder 2 is returned to the reservoir.

Hydraulic Brakes
The hydraulic brake system used in the automobile is a multiple piston system. A multiple piston
system allows forces to be transmitted to two or more pistons in the manner indicated in figure
2.27. Note that the pressure set up by the force applied to the input piston (1) is transmitted
undiminished to both output pistons
(2 and 3), and that the resultant
force on each piston is proportional
to its area. The multiplication of
forces from the input piston to each
output piston is the same as that
explained earlier.

The hydraulic brake system from


the master cylinders to the wheel
cylinders on most automobiles
operates in a way similar to the
system illustrated in figure 2.28.

Figure 2.27: Principle of hydraulic


brake

When the brake pedal is depressed, the pressure on the brake pedal moves the piston within
the master cylinder, forcing the brake fluid from the master cylinder through the tubing and
flexible hose to the wheel cylinders. The wheel cylinders contain two opposed output pistons,
each of which is attached to a brake shoe fitted inside the brake drum.

Module 2.2 Mechanics 2-49


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Figure 2.28: Hydraulic brake system

Each output piston pushes the attached brake shoe against the wall of the brake drum, thus
retarding the rotation of the wheel. When pressure on the pedal is released, the springs on the
brake shoes return the wheel cylinder pistons to their released positions. This action forces the
displaced brake fluid back through the flexible hose and tubing to the master cylinder.

The force applied to the brake pedal produces a proportional force on each of the output
pistons, which in turn apply the brake shoes frictionally to the turning wheels to retard rotation.

As previously mentioned, the hydraulic brake system on most automobiles operates in a similar
way, as shown in figure 2.28. It is beyond the scope of this manual to discuss the various brake
systems.

Accumulators
An accumulator is a pressure storage reservoir in which hydraulic fluid is stored under pressure
from an external source. The storage of fluid under pressure serves several purposes in
hydraulic systems.

In some aircraft hydraulic systems it is necessary to maintain the system pressure within a
specific pressure range for long periods of time. It is very difficult to maintain a closed system
without some leakage, either external or internal. Even a small leak can cause a decrease in
pressure. By using an accumulator, leakage can be compensated for and the system pressure
can be maintained within an acceptable range for long periods of time.

Accumulators also damp out fluctuations in pressure due to the operation of services such as
control surfaces and landing gear. They can supply extra pressure when all the hydraulic
services are being operated at one time (flaps, control surfaces, landing gear etc.) and when the
hydraulic pump is unable to cope. They can also be used in an emergency when all other
hydraulic power pressure supplies (pumps etc) have failed. Thus a large modern aircraft can be
controlled on accumulator power alone, for up to an hour.

Accumulators also compensate for thermal expansion and contraction of the liquid due to
variations in temperature.

2-50 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Figure 2.29: Hydraulic Accumulator

The accumulator consists of an air chamber, which is charged with air or nitrogen. This is
called the pre-charge pressure and is usually about 1000 PSI. This pressure is measured when
there is no hydraulic pressure. The air chamber is the under side of the piston shown in figure
2.29. With no hydraulic pressure, the air/nitrogen pressure will push the piston to the top of the
accumulator. A pressure gauge may be attached to the accumulator to indicate the air/nitrogen
pressure. When the hydraulic pumps are switched on, the hydraulic pressure (acting on top of
the piston, in opposition to the air/nitrogen pressure) begins to rise. When the hydraulic

Module 2.2 Mechanics 2-51


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
pressure exceeds the air/nitrogen pre-charge pressure (1000 PSI), the piston will begin to move
down and further compress the air/nitrogen pressure.

At all times that the hydraulic pressure is above the air/nitrogen pre-charge pressure of 1000
PSI, the air/nitrogen and the hydraulic pressures are equal. Thus when the hydraulic pressure
has reached its working level of 3000 PSI, the air/nitrogen pressure is also 3000 PSI.

It is the additional pressure supplied to the air/nitrogen by the hydraulic pressure, which can be
used to feed back the pressure to the hydraulic fluid if the hydraulic fluid pressure falls below
that of the air/nitrogen. However, when the air/nitrogen gauge indicates 1000 PSI, the hydraulic
pressure is zero, since the air/nitrogen has expanded back to its original pre-charge pressure.

2-52 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Problems

1. Calculate the pressure on a gas when a force of 3100 N is exerted on a piston of


diameter 2 cm

2. Calculate the force exerted when a pressure of 1 bar acts on a piston of diameter
8 cm which has a piston rod of diameter 2 cm taking some of the piston area.

3.

The piston face area in the hydraulic jack shown above is 0.3 sq.in. The rod cross
sectional area is 0.1 sq.in. Calculate the force and direction the ram rod will move if
a pressure of 12 PSI enters equally into both sides of the cylinder chamber.

4. A brake master cylinder has a piston diameter of 0.4 ins. It feeds pressure to 4
identical wheel cylinders, each having just one piston of diameter 2 ins. What is the
force on one wheel brake when the driver applies a force of 80 lbs to the master
cylinder?

5. An hydraulic accumulator is charged with nitrogen to 600 PSI. The hydraulic pump is
then switched on and it feeds 3000 PSI to the other side of the accumulator piston.
What will be the new pressure on the nitrogen side of the accumulator?

Module 2.2 Mechanics 2-53


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-54 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Answers

1. 10 MPa
2. 450 N
3. 1.2 lbs, right
4. 2000 lbs
5. 3000 PSI

Module 2.2 Mechanics 2-55


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-56 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Barometers
A barometer is an instrument used to measure atmospheric pressure. It can measure the
pressure exerted by the atmosphere by using water, air, or mercury.

Mercury barometers
A standard mercury barometer has a glass tube of
about 30 inches (about 76 cm) in height, closed at
one end, with an open mercury-filled reservoir at
the base. Mercury in the tube adjusts until the
weight of the mercury column balances the
atmospheric force exerted on the reservoir. High
atmospheric pressure places more force on the
reservoir, forcing mercury higher in the column.
Low pressure allows the mercury to drop to a
lower level in the column by lowering the force
placed on the reservoir.

Since higher temperature at the instrument will


reduce the density of the mercury the scale for
reading the height of the mercury is adjusted to
compensate for this effect.

The standard temperature for reading a mercury


barometer is 0oC (32oF). A correction factor is
read from a graph and applied to the reading Figure 2.30: Mercury Barometer
for temperatures above 0oC. The barometer
over-reads at higher temperatures.

The mercury barometer's design gives rise to the expression of atmospheric pressure in inches
or millimetres: the pressure is quoted as the level of the mercury's height in the vertical column.
1 atmosphere is equivalent to about 29.9 inches, or 760 millimetres, of mercury. Barometers of
this type normally measure atmospheric pressures between 28 and 31 inches of mercury.

The reading from a barometer (in mm.hg or in.hg) can be converted into Pascals or PSI by
using the formula

Pressure = ρgh (if using Metric units)

Or

Pressure = ρh (if using English units)

Where:

Pressure = the converted pressure (Pa or PSI)


ρ = density of the mercury (13,600 kg/m3 or 62.4 lbf../ft3)
h = height of mercury (m or ft)

Module 2.2 Mechanics 2-57


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Aneroid barometers

An aneroid barometer uses a small, flexible metal


vacuum chamber called an aneroid cell. This aneroid
capsule (cell) is made from an alloy of beryllium and
copper. The evacuated capsule (or usually more
capsules) is prevented from collapsing by a strong
spring. Small changes in external air pressure cause
the cell to expand or contract. This expansion and
contraction drives mechanical levers such that the tiny
movements of the capsule are amplified and displayed
on the face of the aneroid barometer. Many models
include a manually set needle which is used to mark
the current measurement so a change can be seen. In
addition, the mechanism is made deliberately 'stiff' so
that tapping the barometer reveals whether the
pressure is rising or falling as the pointer moves. They
Figure 2.31: Aneroid barometer
are used for measuring atmospheric pressure.

2-58 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Buoyancy

Archimedes Principle
Archimedes was a Greek philosopher and mathematician who lived about 250BC. There is a
story (maybe even true) about Archimedes that every physics student should hear. It goes as
follows:

The king who ruled Greece at that time asked his royal metalworkers to make him a gold
crown. When the crown was delivered it was indeed beautiful. However, the king suspected
that the crown was not pure gold. He did not want to destroy the crown but he wanted to know
if he had been cheated. What he needed was some type of non-destructive evaluation (NDE
dates back many years!). He asked Archimedes to solve his problem. Archimedes pondered
the question. The density (mass/volume) of gold was well known. He knew of course how to
determine the weight and mass of the crown by simple weighing. However, since the crown
did not have a regular shape it was impossible to determine the volume by a mathematical
calculation. The solution came to Archimedes one day when his servant filled his bathtub too
full. As Archimedes stepped into his bath, he noticed that a volume of water equal to his
volume overflowed! With a flash of insight he ran through Athens, stark naked, shouting
“Eureka, Eureka, I have the solution!” The experiment was performed, the king was notified
that his crown was not pure gold and the royal metal workers lost their lives.

The point of the above story is that a body submerged in a liquid displaces a volume of water
equal to its own volume. A corollary is that a body that floats in a liquid displaces a volume of
liquid less than its volume since some portion of the body is above the water level.

Archimedes' principle, states that a body


immersed in a fluid is buoyed up by a force equal to
the weight of the displaced fluid. The principle
applies to both floating and submerged bodies and
to all fluids, i.e., liquids and gases. It explains not
only the buoyancy of ships and other vessels in
water but also the rise of a balloon in the air and the
apparent loss of weight of objects underwater.

In determining whether a given body will float in a


given fluid, both weight and volume must be
considered; that is, the relative density, or weight
per unit of volume, of the body compared to the fluid
determines the buoyant force. If the body is less
dense than the fluid, it will float or, in the case of a
balloon, it will rise. If the body is denser than the
fluid, it will sink.
Figure 2.32: Archimedes’ principle

Relative density also determines the proportion of a floating body that will be submerged in a
fluid. If the body is two thirds as dense as the fluid, then two thirds of its volume will be

Module 2.2 Mechanics 2-59


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
submerged, displacing in the process a volume of fluid whose weight is equal to the entire
weight of the body.

In the case of a submerged body, the apparent weight of the body is equal to its weight in air
less the weight of an equal volume of fluid. The fluid most often encountered in applications of
Archimedes' principle is water, and the specific gravity of a substance is a convenient
measure of its relative density compared to water.

In calculating the buoyant force on a body, however, one must also take into account the shape
and position of the body. A steel rowboat placed on end into the water will sink because the
density of steel is much greater than that of water. However, in its normal, keel-down position,
the effective volume of the boat includes all the air inside it, so that its average density is then
less than that of water, and as a result it will float.

Archimedes’ Principle Applied to Bodies that Float


A body will float in any liquid that has a weight density greater than the weight density of the
body. For example a body of weight density 63.4 lbs./ft.3 would float in ocean water (D = 64.4
lbs./ft.3) and sink in lake water (D = 62.4 lbs./ft.3).

When bodies float they can float “high” or float “low”. The ratio of the weight density of the
floating body relative to the weight density of the liquid determines exactly how high or low a
body will float.

In order to understand Archimedes’ Principle as applied to floating bodies, let us consider a


submarine and imagine that a block of wood of weight density 48.3 lbs./ft.3 and volume 2 ft.3 is
thrust out of the hatch of a submarine into the ocean water. We know intuitively that this block of
wood will rise to the ocean surface.

The weight of the block is (48.3 lbs./ft.3) (2 ft.3) = 96.6 lbs. As long as the block is below the
water surface (while it is rising to the top), it displaces 2 ft.3 of ocean water.

We know that:

BF = weight of displaced ocean water

= (64.4 lbs./ft.3) (2 ft.3)

BF = 128.8 lbs.

Figure 2.33: Archimedes’ principle

We can see why the block rises. How far will the block rise? it will rise until the BF exactly
equals its weight. In our example it will rise until the BF has been reduced to 96.6 lbs. (the
weight of the block). The BF will be reduced as the block emerges from the water. In our
example, it will rise until 25% of the block’s volume is above the water surface. It follows that
75% of 2 ft.3 (= 1.5 ft.3 ) will be below the water surface. When this occurs, the BF on the block
is (64.4 lbs./ft.3)(1.5 ft.3) equals 96.6 lbs. Note again that the BF equals the weight of the block
while the block is floating.

2-60 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
In the preceding example, note that the ratio of the weight density of the block (48.3 lbs./ft.3) to
the weight density of the ocean water (64.4 lbs./ft.3) was 0.75. We recall that 75% of the floating
block was under water. This is generally true and makes a much easier procedure to determine
how low a block will float in a given liquid.

In dealing with bodies that float, it is important to


note that boats, made of materials more dense
than water, are shaped in such a way that the total
weight density is less than water. In order to
understand this, consider the rowing boat with con-
tents (people, lunch, fishing gear, etc.) shown in
figure 2.34. Note that some of the boat (shown with
dotted lines) is below the water surface. Suppose
that the row boat floats in such a way that it
displaces 8 cu ft. of lake water. The weight of the
displaced water is 8 ft. (62.4 lbs./ft.3) or 499 lbs.
Figure 2.34 Therefore, the BF is 499 lbs. The boat and
contents must weigh 499 lbs. to float at this level. If
the boat weighs 150 lbs. the contents must weigh 349 lbs.

This is realistic (father 200 lbs., son 75 lbs., lunch 25 lbs., fishing gear 49 lbs.).

One final comment should be made regarding submarines. Submarines cruising at a definite
depth in ocean water have a total weight density equal to the weight density of ocean water,
64.4 lbs./ft.3 This means that the total weight of the submarine (metal shell, air, crew, load,
ballast, etc.) divided by the total volume is 64.4 lbs./ft.3 The ballast used in submarines is
ocean water. These vessels can take on water or pump out water. If the submarine wants to
descend, it takes on water. If it wants to rise toward the surface it pumps out water.

EXAMPLE:

A block of oak (D = 45 lbs./ft.3) is placed in a tank of benzene (D = 54.9 lbs /ft.3). The oak
floats since its weight density is less that the weight density of the benzene. What percentage
of the oak will be below the surface of the benzene?

We find the ratio of the two weight densities.

45 lbs. / ft 3
= 0 .82
54 .9 lbs / ft 3

We conclude that 82% of the oak block will be below the surface of the benzene.

Module 2.2 Mechanics 2-61


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Archimedes’ Principles as Applied to Airships and Balloons
In all of the above materials, we have talked about Archimedes’ principle as if it applied only to
liquids. Since most of our experience with this principle is with liquids, it seemed easier to do
this at first. However, it must now be emphasized that buoyant forces exist also with gases.

The obvious example is that of a hot air balloon or a lighter-than-air aircraft.

EXAMPLE:

The bag of a balloon is a sphere of radius 25 m filled with hydrogen of weight density 0.882
N/rn3. What total weight (in Newtons) of fabric, car, and contents can be lifted by this balloon
in air of weight density 12.6 N/rn3?

We first calculate the volume of the spherical balloon by recalling that the volume of a sphere
is given by:

4 4
V= π R 3 = (3.1416)(25 m) 3 = 65,450 m 3
3 3

The weight of the hydrogen is found from the formula D V = w:

(0.882 N/m3) (65,450 m3) = 57,700 N.

The weight of the displaced air is:

(12.6 N/rn3) (65,450 m3) = 824,700 N.

Since the weight of the displaced air is the BF we can say that:

BF = 824,700 N

This BF must hold up the hydrogen, fabric, car, and contents. It follows that fabric, car, and
contents weighing 767,000 N can be lifted by this balloon. Note that this number was obtained
by subtracting 57,700 N from 824,700 N.

Usually balloons are not filled with hydrogen since hydrogen is explosive. Of course, since
hydrogen is the lightest of all gases it is the most efficient. However, the danger of explosion
outweighs this advantage. The next lightest gas is helium of weight density 1.74 N/rn3.
Usually, balloons are filled with this gas.

2-62 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Table 2.3: Weight densities for common
Module 2.2 Mechanics 2-63
Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-64 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Problems

1. A solid aluminium object of volume 250 ft3 is resting on the ocean floor. A salvage crew
plans to raise this object. What force will be needed?

2. A solid steel body of volume 125 ft3 is to be raised by a salvaging crew to the surface of a
lake. What force will be needed?

3. What percentage of an iceberg is below the surface of the ocean?

4. A canoe is floating in such a way that it displaces 6 cu.ft. of lake water. If the canoe
weighs 100 lbs., what is the weight of its contents?

5. A balloon is spherical in shape and has a radius of 20 ft. It is filled with helium (weight
density 0.01 lb/ft3) and is floating in air (weight density 0.08 lb/ft3). What is the weight of
the balloon (fabric, crew and contents etc.)?

Module 2.2 Mechanics 2-65


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-66 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Answers

1. 26,200 lbs.
2. 53,100 lbs
3. 89%
4. 274 lbs
5. 2240 lbs

Module 2.2 Mechanics 2-67


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-68 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Kinetics
Linear Motion
When a body is moving in a straight line with constant speed it is not accelerating. We say, in
this case, that it is moving with constant velocity. If a body’s velocity is not constant, it is
accelerating. A body accelerates if it is changing its speed and/or its direction.

When we discuss a body’s straight-line motion, then we do not have any change in direction. In
this special case, any acceleration is due to a change in speed.

The Equations of Motion


In all of the following discussion, certain symbols will be used. These symbols are summarized
below:

Vav = average velocity


t = time
u = initial velocity
v = final velocity
a = acceleration
* s = distance covered

* Note that ‘s’ is the traditional notation for distance in almost all physics textbooks. This
choice reduces confusion with the symbol d for derivative, a concept from calculus.

There is a formula dealing with the motion of a body that you have used for many years. In
school, you probably memorized the formula in these words:

distance = rate (or speed) x time

Using our above symbols, we could write:

(1) s = Vavt

Note that for the rate, we have used the average speed. We all know that even though
sometimes speed changes, we can always talk about the average speed. Thus, if we travel at
an average speed of 50 MPH for 6 hours, we cover 300 miles.

Now we must extend our treatment of motion to include the concept of acceleration.
Acceleration (for straight-line motion) is the rate of change of speed in time. We define
acceleration (for straight-line motion) in the following manner:

v −u
(2) a=
t

In using this formula, a may be either positive or negative. If v is less than u, then our value of a
turns out to be a negative number.

Module 2.2 Mechanics 2-69


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
EXAMPLE:

A truck is initially travelling at a speed of 50 ft./sec. The driver applies his brakes for 15 sec.
The final speed of the car is 20 ft./sec. What is the acceleration?

20 ft / sec − 50 ft / sec
a=
15 sec

− 30 ft / sec
a=
15 sec

− 2 ft / sec
a= = -2 ft/sec/sec
sec

a = -2 ft/sec2

Notice that the unit of acceleration has the square of a time unit in its denominator.

A little thought will convince you that an acceleration is positive if the body is increasing speed
and negative when the body is decreasing its speed. If we cross-multiply in formula (2) we
obtain:
at = v - u

After transposing, we can write:

(3) v = u + at

If an automobile is on an expressway and the driver is increasing speed smoothly and regularly,
we note that his average speed is the average of his initial and final speed.

The equation can be written:

u+v
Vav =
2

If this value of Vav is substituted into equation (1), we have:

u+v
(4) s= t
2

In this equation, we can substitute for v (= u + at) using the value in equation (3).

u + (u + at) 2u + at
s= t = t
2 2

2-70 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
After a bit of algebra, we obtain:

(5) s = ut + ½ at 2

Equation (4) can be written, after cross-multiplication:

2s = (u + v)t

We can now multiply this equation by equation (2). After cancelling time (t) on the right:

2as = (v - u) (v + u)
or 2as = v2 – u2

The final form of this formula is:

(6) v2 = u2 + 2as

These equations are very important. They enable us to deal with all kinds of motion problems
where the body is in straight line motion and is changing its speed. These formulas will be
summarized below. They will be numbered with Roman numerals and can be referred to by
these numbers when used in the problem exercises.

u+v
i. s= t
2

ii. v = u + at

iii. s = ut + ½ at 2

iv. v2 = u2 + 2as

When a body in straight line motion is not changing speed, or in cases where we are interested
only in the average speed, the formula is more simple.

s = Vavt

Formulas i through iv are used in many practical physics problems. Note that each one involves
four quantities. When a problem is given to you to solve, be sure to determine which of these
three quantities are given to you, and which quantity is to be found. Choose the formula which
involves these four quantities. If the formula is not solved for the unknown quantity, solve for this
quantity algebraically. Finally substitute the known quantities and solve for the unknown
quantity.

An example should clarify the above procedure.

Module 2.2 Mechanics 2-71


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
EXAMPLE:

An automobile has an initial speed of 50 ft./sec. and a final speed of 75 ft./sec. While it is
undergoing this change of speed, it travels a distance of 125 ft. What is its acceleration?

In attacking this problem it is wise to write down exactly what is known and what is unknown.
u = 50 ft./sec. V = 75 ft/sec.
S = 125ft. a=?

Formula iv involves these four quantities. Note that i, ii, and ill do not involve these exact four
quantities. Formula iv is the one to use. First it should be solved for the unknown, a.

v2 = u2 + 2as

v2 - u2 = 2as

v 2 − u2
a=
2s

(75 ft / sec) 2 − (50 ft / sec) 2


a=
2 (125ft )

3125 ft 2 / sec 2 ft 2 1
a= = 12 .5
250 ft sec 2 ft

a = 12.5 ft/sec2

Accelerated Motion of a “Freely Falling” Body


Common experience indicates that falling bodies accelerate or increase in speed as they fall.
Close to the surface of the earth this “acceleration of a freely falling body” has been measured
to be about 32 ft./sec.2 in the English system and 9.8 m/sec.2 in the metric system. The “about”
in the preceding sentence indicates that this quantity varies somewhat over the face of our
earth. The values given are average values.

When we use the words “freely falling”, we mean that we are neglecting the effects of air
resistance (as if we were in a vacuum). Of course, there is always air resistance, so how can we
neglect it?

When a body is falling with a great speed, air resistance can certainly not be neglected. To use
the acceleration formulas in these cases would give us results that are not valid. However, if a
body is falling close to the surface of the earth, the acceleration formulas do give us valid results
if the height from which it falls is not too great.

2-72 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Some numerical data should clarify the preceding statements. If a compact body, such as a
stone, is dropped (not thrown) from a height of 324 ft. above the surface of the earth, it will take
about 4.5 sec. for the body to reach the ground. It will have obtained a speed of 144 ft/sec. (98
MPH). At this speed, the effects of air resistance are still quite negligible. Above this speed (98
MPH), the effects of air resistance are not negligible.

Therefore, we can conclude that the fall of a body from a height of 324 ft. or less (or equivalently
during a time of 4.5 sec. or less) can be handled quite accurately with the ordinary acceleration
formulas. The value of the acceleration will be either 9.8 m/sec2 or 32 ft./sec2 if the body is rising
and therefore decreasing its speed the values of the acceleration will be - 9.8 m/sec2 or - 32
ft/sec2

If a body falls from a height greater than about 324 ft. above the surface of the earth, the air re-
sistance becomes very important. As we have said, a height of 324 ft. corresponds to a fall of
4.5 sec. When the time of fall increases to about 8 seconds, the speed of fall has increased to
about 115 MPH. When the time of fall is between 4.5 sec. and 8 sec. the speed increases in a
non-linear manner from 98 MPH to 115 MPH. As the time of fall increases beyond 8 seconds
the speed of fall remains constant at about 115 MPH. This speed of fall is called the “terminal
velocity”.

All of the above data indicates that it is possible to use the acceleration formulas with accurate
results for many applications dealing with falling bodies. We will limit our applications to cases
where the formulas are valid: heights less than 324 ft. and times of fall less than 4.5 seconds.

EXAMPLE

A body started from rest and has been falling freely for 3 sec. At what speed is it falling?

u = o, t = 3 sec, a = 32 ft/sec2, v = ?

We will use Formula ii.

v = u + at

v = 0 +  32
ft 
(3 sec)
 sec 2 

v = 96 ft/sec

Module 2.2 Mechanics 2-73


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
EXAMPLE

A body started at rest and has been falling freely for 3 sec. How far has it fallen?

u = 0, t = 3 sec, a = 32 ft/sec2, s = ?

We will use Formula iii.

s = ut + ½ at 2

1 ft 
 32 ( 3 sec)
2
s = (0)(3 sec) +
2 sec 2 

s = 144 ft

EXAMPLE:

A body is thrown upward with an initial speed of 120 ft./sec. How high does it rise?

u = 120 ft/sec, v = 0, a = -32 ft/sec2, s = ?

We will use Formula iv.


v2 = u2 + 2as

V 2 − u2
s=
2a

0 − (120 ft / sec) 2
s=
2 ( −32 ft / sec 2 )

ft 2 sec 2
s = 225
sec 2 ft

s = 225 ft.

2-74 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Problems

1. A car on the motorway is accelerating at 25ft/sec2. If it started from rest and has been
accelerating for 5 sec., how far has it travelled during this time of acceleration?

2. A truck had an initial velocity of 40ft/sec. It accelerated at 10ft/sec2 and reached a final
velocity of 60ft/sec. How far did this truck travel while it was accelerating?

3. A car slowed down from 80 ft/sec. to 40 ft/sec. while travelling a distance of 100 ft. What
was its acceleration?

4. A car, originally travelling at 25 ft/sec, increases its speed at a rate of 5 ft/sec2 for a
period of 6 sec. What was its final speed?

5. A car has an initial velocity of 40 ft/sec. It slows down at a rate of 5 ft/sec2 and covers a
distance of 60 ft. while slowing down. What is its final velocity?

6. A stone is dropped from a high building and falls freely for 4 sec. How far (in meters) has
it fallen during this time?

7. A stone is thrown upward with an initial velocity of 64 ft/sec. How high does it rise?

8. A ball is dropped from a bridge into the river below and 2.5 sec. after the ball is dropped
a splash is heard in the water below. How high is the bridge?

9. A car starts with an initial velocity of 30 ft/sec. and accelerates for 5 sec. at
4 ft/sec2. How far has it travelled during this time?

10. A Cessna Agcarryall has a take-off run of 900 feet, at the end of which its speed is 80
MPH. How much time does the run take?

(Hint: convert MPH to ft./sec. first)

11. A Grumman Tomcat, powered by two Pratt & Whitney turbofan engines, has a maximum
acceleration during take-off of 20 ft/sec2. What velocity can it achieve by the end of a
1000 foot take-off run?

Module 2.2 Mechanics 2-75


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-76 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Answers

1. 312 ft.

2. 100 ft.

3. -24 ft/sec2

4. 55ft/sec.

5. 32 ft/sec.

6. 78 m

7. 64 ft.

8. 100 ft. or 30.6m

9. 200 ft.

10. 15 sec.

11. 200 ft/sec

Module 2.2 Mechanics 2-77


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-78 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Rotational Motion

Introduction
Previously we discussed constant speed and accelerated motion in a straight line and derived
four important formulas which will be reviewed below. In this chapter we will consider motion
which takes place on a circular path. Such motion is very common in our complex society and
we need to understand more about motion in curved paths.

Degrees and Radians


Before we begin our discussion, we need to define a new unit for measuring angles, the radian
(see figure 2.35).

A radian is defined as the central angle subtending a length of arc equal to the radius of the
circle.

A radian is approximately equal to 57.3°. The conve rsion factors for angle units are:

1 revolution = 360°
1 revolution = 2π radians
2π radians = 360°
1 radian = 57.3°

Now let us consider a body (represented by a point)


moving in a circular path. An initial reference line is
shown in figure 2.36. As the point moves about the
circle in a counter-clockwise sense, a line drawn
between the point and the centre of the circle
continuously sweeps out an angle. This angle can be
measured in revolutions, radians or degrees. We call
this angle the angular displacement of the point and
use the Greek letter theta (θ) to represent this angular
displacement.
Figure 2.35
If the point moves with constant speed it also has a constant angular velocity. That is, the line
drawn from the point to the centre of the circle sweeps out a definite number of revolutions,
radians, or degrees each second or minute. The symbol used to represent angular velocity is
the Greek letter omega (ω).

Angular velocity can be expressed in different units, such as,

radians rev. degrees


sec . sec . sec.

radians rev. degrees


min. min .. min.

Module 2.2 Mechanics 2-79


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
It is also possible that the point is not moving with constant angular velocity. It may be
increasing or decreasing its angular velocity. When a CD starts rotating in a CD drive the
angular velocity increases until it reaches a constant value. After the reject button is pushed the
angular velocity decreases until the CD comes to rest.

In both of the above cases we say that the point has an angular acceleration. The Greek letter
alpha (α) is used for angular acceleration. Note that α is positive if the angular velocity is
increasing and negative if the angular velocity is decreasing.

Angular acceleration can also be expressed in different units,

radians rev. degrees


sec . 2 sec . 2 sec.2

radians rev. degrees


min . 2 min.2 min.2

Now as a body moves in a circular path four similar


equations hold as in the case of a body moving in a
straight-line path. Both sets of equations will be shown
below. It is important to re-memorize the equations for
straight-line motion. In this way the other four equations will
also be known, since they are exactly analogous.

Figure 2.36: A point moving in a circle

u+v ω1 + ω 2
s= t θ= t
2 2
v = u + at ω 2 = ω1 + αt
s = ut + ½ at 2 θ = ω1t + ½ αt2
v2 = u2 +2as ω22 = ω12 +2αθ

Study these equations carefully and note that the set to the right, the “rotational analogy” are
easily remembered if the left set is well known. We recall that the subscripts ”u” and ”v” indicate
“initial” and “final”.

These four rotational equations help us to solve many practical problems dealing with rotating
bodies.

2-80 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
EXAMPLE:

A rotating machine part increases in angular velocity from 3 rev./min. to 35 rev./min. In 3.5
minutes. What is its angular acceleration?

We use the following equation and solve it for a.

ω 2 = ω1 + αt
ω 2 − ω1

t
We now substitute our known values.

35 rev / min − 3 rev / min


α= = 9.14 rev / min 2
3.5 min

EXAMPLE:

A propeller starts from an angular velocity of 900 rev./min. and accelerates at 100
rev./min.2 for 5 minutes. Through how many revolutions has it turned?

θ = ω1t + ½ αt2

θ = (900 rev./min.)(5 min) + ½ (100 rev./min. 2)(5 min.)2

θ = 5,750 revolutions

EXAMPLE:

A propeller starts at 1,000 rev./min. and accelerates at 100 rev./min.2 through 2,000
revolutions. What is its final angular velocity?

ω22 = ω12 +2αθ

ω22 = (1,000 rev./min.)2 +2 (100 rev./min.2)(2,000 rev.)

ω2 = 1,180 rev./min.

Module 2.2 Mechanics 2-81


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
We note that there is an acceleration of the body “in the
path”, called the tangential acceleration. The body is
increasing or decreasing its speed, or traversing the circle.
We recall also that when a body moves in a circle there is
also a centripetal acceleration, V2/R, that is always
directed toward the centre of the circular path.

Thus when a body is increasing speed as it moves in a


circular path there are two acceleration vectors, one
tangential to the path, and the other directed to the centre
of the path (centripetal acceleration). In figure 2.37, the
body is increasing speed in the counter-clockwise sense.
The directions of the two acceleration vectors are shown.

Figure 2.37: Tangential


acceleration (at) and Centripetal
acceleration (ac)
Radian Measure
In figure 2.38, ‘s’ is the length along the path. We would like to relate this distance to the size of
the central angle (θ) and the radius (R) of the circular path. In our preceding discussion, the
angle (θ) was measured in any of three different units, degrees, revolutions, or radians.

Figure 2.38: s, R and θ

The equation that relates s to θ and R is a very simple one if we limit the angular unit to radians.
This equation is:
S = Rθ

We see that this equation is true if we look at figure 2.38. We note, by measuring, that the
equation is satisfied. We also see that it would not be true if the angle θ was in revolutions or
degrees

We now have a new problem to deal with in our treatment of rotational motion. There is a limit to
the units that may be used in this equation. We repeat that, for this equation, we must use
radian measure. Also, any equation that is derived from s = Rθ will have this same restriction.

2-82 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Suppose that a body moves a small distance along the path and sweeps out a small central
angle.

The usual mathematical notation for a very small quantity is the use of the Greek letter Delta
(∆).

∆s = R ∆θ

Let us divide both sides of this equation by the time, (∆t) during which the motion occurred.

∆s ∆θ
=R
∆t ∆t
We can write:

ω
v = Rω

If this velocity in the path is changing. there is also a change in the angular velocity. Assume
that this change occurs in the small time interval (∆t).

We can write:
∆v = R ∆ω

Next we divide left and right members by ∆t.

∆v ∆ω
=R
∆t ∆t

The tangential acceleration (a) in the left side is the rate at which a body moving in a circular
path is picking up speed in the path. It is equal to the radius times the angular acceleration (α).

We can write:
α
a = Rα

Let us summarize the three important equations we have derived:

θ
s = Rθ
ω
v = Rω
α
a = Rα

All three of these equations require the use of radian measure. This means that:

θ must be in radians
ω must be in rad/min. or rad./sec.
α must be in rad./min2 or rad./sec2

Note that the radian is called a “dimensionless” unit. We put it in or take it out for clarity.

Module 2.2 Mechanics 2-83


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
EXAMPLES:

A car is moving on a circular racetrack of radius 150 ft. It sweeps out an angle of 2000.
How far has it travelled?

We note that:
6 .28 rads .
θ = 200° x = 3.49 rads.
360 o

s = Rθ

s = (150 ft.) (3.49 rad.)

s = 523 ft.

3.36 rev./min.

EXAMPLE:

A race car is travelling at a speed of 176 ft./sec. (120 MPH) around a circular racetrack of
radius 500 ft. What is the angular velocity of this car in rev./min.?

Use the equation:


v = Rω

v 176 ft. / sec .


or ω= =
R 500 ft

ω = 0.352 rad./sec.

Note that we knew that the unit of our answer is rad./sec. and not rev./sec. since the
equation we used always is in radian measure. The units in the right side of the second
equation above actually come out as “nothing”/sec. We put in the radian unit in the
numerator for clarity.

in order to find our answer in rev./min. we use the proper conversion factors.

0 .352 rad. / sec . × 60 sec .


ω= = 3.36 rev./min.
6 .28 rad.

2-84 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
EXAMPLE:

A race car is moving on a circular racetrack of radius 4,000 ft.. It is increasing its speed at
a rate of 15 ft./sec.2 What is its angular acceleration rev./sec.2?

We use the equation:

a = Rα

a 15 ft / sec 2
α= =
R 4,000 ft

α = 0.00375 rad./sec.2

We note that the unit is rad./sec.2 because the equation that we have used requires radian
measure.

To obtain a in rev./sec.2, we must use the standard conversion factor.

0.00375 rad / sec 2


α=
6 .28 rad.

α = 0.000597 rev./sec.2

Module 2.2 Mechanics 2-85


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-86 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Problems

1. A propeller starts from rest and accelerates at 120 rev/sec2 for 4 seconds. What is its
final angular velocity in rev/sec? In rev/min?

2. A rotating turntable starts from rest and accelerates at 5 rev/min2 for 3 min. Through how
many revolutions has it turned?

3. A helicopter main rotor starts from an initial angular velocity of 2 rev/min and accelerates
at 60 rev/min2 while turning through 400 revolutions. What is its final angular velocity?

4. A plane is circling O’Hare in a circular pattern of radius 15,000 ft. It sweeps out an angle
of 340o? How far has it travelled?

5. A plane is circling an airport in a circle of radius 5,000 ft. How far has it travelled after 4
revolutions?

6. A race car is moving on a circular track of radius 600 ft. It is travelling at a speed of 100
ft/s. What is its angular velocity in rev/min?

7. A race car is moving on a circular racetrack of radius 800 ft. It is accelerating at a rate of
10 ft/sec2 What is its angular acceleration in rev/sec2?

8. A helicopter tail rotor starts with an initial angular velocity of 15 rev/sec and decelerates
at a rate of 2.00 rev/sec2 until it comes to rest. Through how many revolutions has the
rotor turned while it comes to rest?

Module 2.2 Mechanics 2-87


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-88 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Answers

1. 480 rev/sec. 28,800 rev/min.


2. 22.5 rev.
3. 219 rev/min.
4. 89,000 ft.
5. 23.8 miles
6. 1/6 rad/s, 5/π rev/min
7. 1/160π rev/sec2
8. 56.3 rev.

Module 2.2 Mechanics 2-89


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-90 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Periodic Motion

Simple Pendulum
A simple pendulum is one which can be considered to be a point mass
suspended from a string or rod of negligible mass. It is a resonant system with
a single resonant frequency. For small amplitudes, the period of such a
pendulum can be approximated by:

Where: L = the length of the pendulum is m, or ft


g = the magnitude of acceleration due to gravity = 9.81
m/s2 or 32 ft/s2

Note: The Natural Frequency of Oscillation is independent of the mass of the


pendulum, and of the amount of initial displacement

Figure 2.39: A simple


pendulum

This expression for period is reasonably accurate for angles of a few degrees, but the treatment
of the large amplitude pendulum is much more complex

It is interesting to note that the pendulum will oscillate at only one frequency, regardless of how
far the pendulum is initially displaced, or for how long the pendulum is left to oscillate. The only
factor that changes, is the linear velocity of the mass. This fixed frequency is known as the
Natural Frequency of Oscillation.

If we consider only the horizontal motion of the mass and neglect its vertical motion as it swings
(an assumption which can be made if the string is long compared to the amplitude of swing),
then the periodic motion is said to be Simple Harmonic Motion (SHM).

Time period (T) and frequency (f) can also be related to each other by the formulae:

1 1
T= or f=
f T

Module 2.2 Mechanics 2-91


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Mass and Spring
When a mass is acted upon by an elastic force which tends to bring it back to its equilibrium
position, and when that force is proportional to the distance from equilibrium (e.g., doubles
when the distance from equilibrium doubles - a Hooke's Law force), then the object will undergo
periodic motion when released.
A mass on a spring is the standard
example of such periodic motion. If
the displacement of the mass is
plotted as a function of time, it will
trace out a pure sine wave.

The motion of the medium in a


travelling wave is also simple
harmonic motion as the wave
passes a given point in the medium.

Figure 2.40: Sinusoidal motion of a spring/mass system

It is interesting to note that the spring/mass system will oscillate at


only one frequency, regardless of how far the mass is initially displaced, or for how long the
system is left to oscillate. The only factor that changes, is the linear velocity of the mass. The
fixed frequency is known as the Natural Frequency of Oscillation, and and can be calculated from
the formula:

Where: k = the stiffness of the spring in N/m, or lb/in


m = the mass of the oscillating body

Note: The Natural Frequency of Oscillation is independent of the magnitude of gravity, and of
the amount of initial displacement

Simple Harmonic Motion (SHM)

What is SHM
Motion which repeats itself precisely and can be described with the following terms:

• Period: the time required to complete a full cycle, T in seconds.


• Frequency: the number of cycles per second, f in Hertz (Hz)
• Amplitude: the maximum displacement from equilibrium, A

and if the periodic motion is in the form of a travelling wave, one needs also

• Velocity of propagation: v
• Wavelength: repeat distance of wave, λ.

2-92 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Simple harmonic motion is the motion of a simple harmonic oscillator (such as a pendulum or
spring/mass system), a motion that is neither driven nor damped. The motion is periodic, as it
repeats itself at standard intervals in a specific manner - described as being sinusoidal, with
constant amplitude. It is characterized by its amplitude, its period which is the time for a single
oscillation, its frequency which is the number of cycles per second, and its phase, which
determines the starting point on the sine wave. The period, and its inverse the frequency, are
constants determined by the overall system, while the amplitude and phase are determined by
the initial conditions (position and velocity) of that system.

A single frequency travelling wave will take the form of a sine wave. A snapshot of the wave in
space at an instant of time can be used to show the relationship of the wave properties
frequency, wavelength and propagation velocity.

Figure 2.41: The sinusoidal waveform terminology

The motion relationship "distance = velocity x time"


time" is the key to the basic wave relationship.

With the wavelength as distance, this relationship becomes λ=vT. Then using f=1/T gives the
standard wave relationship

This is a general wave relationship which applies to sound and light waves, other
electromagnetic waves, and waves in mechanical media.

Properties of SHM
Considering the motion of a mass on the end of a spring, or the horizontal motion of a
pendulum, the following properties can be observed:

• The velocity of the body is always changing. It is maximum at the undisturbed position
(centre of its motion) and zero at the extremities of its motion (maximum displacement
position)
• The acceleration of the body isis always changing. It is maximum at the extremities of its
motion (maximum displacement position) and zero at its undisturbed position (centre of
motion).

In other words, when its velocity is zero, its acceleration is a maximum, and when its
acceleration is zero, its velocity is a maximum.

Module 2.2 Mechanics 2-93


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Vibration
Vibration refers to mechanical oscillations about an equilibrium point. The oscillations may be
periodic such as the motion of a pendulum or random such as the movement of a tire on a
gravel road.

Vibration is occasionally desirable. For example the motion of a tuning fork, the reed in a
woodwind instrument or harmonica, or the cone of a loudspeaker is desirable vibration,
necessary for the correct functioning of the various devices.

More often, vibration is undesirable, wasting energy and creating unwanted sound -- noise. For
example, the vibrational motions of engines, electric motors, or any mechanical device in
operation are typically unwanted. Such vibrations can be caused by imbalances in the rotating
parts, uneven friction, the meshing of gear teeth, etc. Careful designs usually minimise
unwanted vibrations.

The study of sound and vibration are closely related. Sound, pressure waves, are generated by
vibrating structures (e.g. vocal cords) and pressure waves can generate vibration of structures
(e.g. ear drum). Hence, when trying to reduce noise it is often a problem in trying to reduce
vibration.

Types of vibration

Free vibration occurs when a mechanical system is set off with an initial input and then allowed
to vibrate freely. Examples of this type of vibration are pulling a child back on a swing and then
letting go or hitting a tuning fork and letting it ring. The mechanical system will then vibrate at
one or more of its natural frequencies and damp down to zero.

Forced vibration is when an alternating force or motion is applied to a mechanical system.


Examples of this type of vibration include a shaking washing machine due to an imbalance,
transportation vibration (caused by truck engine, springs, road, etc), or the vibration of a building
during an earthquake. In forced vibration the frequency of the vibration is the frequency of the
force or motion applied, but the magnitude of the vibration is strongly dependent on the
mechanical system itself.

Resonance

What is Resonance?
Resonance is the phenomenon of producing large amplitude of vibrations by a small periodic
driving force. It is the tendency of a system to oscillate at maximum amplitude at a certain
frequency. This frequency is known as the system's resonance frequency (or resonant
frequency). When damping is small, the resonance frequency is approximately equal to the
natural frequency of the system, which is the frequency of free vibrations. Under resonance
condition the energy supplied by the driving force is sufficient enough to overcome friction.

Examples of Resonance
One familiar example is a playground swing, which is a crude pendulum. When pushing
someone in a swing, pushes that are timed with the correct interval between them (the resonant

2-94 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
frequency), will make the swing go higher and higher (maximum amplitude), while attempting to
push the swing at a faster or slower rate will result in much smaller arcs.

Other examples:

• acoustic resonances of musical instruments


• the oscillations of the balance wheel in a mechanical watch
• electrical resonance of tuned circuits in radios that allow individual stations to be picked
up
• the shattering of crystal glasses when exposed to a strong enough sound that causes the
glass to resonate.

A resonator, whether mechanical, acoustic, or electrical, will probably have more than one
resonance frequency (especially harmonics of the strongest resonance). It will be easy to
vibrate at those frequencies, and more difficult to vibrate at other frequencies. It will "pick out" its
resonance frequency from a complex excitation, such as an impulse or a wideband noise
excitation. In effect, it is filtering out all frequencies other than its resonance.

What Causes Resonance?


Resonance is simple to understand if you view the spring and mass as energy storage elements
- the mass storing kinetic energy and the spring storing potential energy. When the mass and
spring have no force acting on them they transfer energy back forth at a rate equal to the
natural frequency. In other words, if energy is to be efficiently pumped into the mass and spring
the energy source needs to feed the energy in at a rate equal to the natural frequency. Applying
a force to the mass and spring is similar to pushing a child on swing - you need to push at the
correct moment if you want the swing to get higher and higher. As in the case of the swing, the
force applied does not necessarily have to be high to get large motions. The pushes just need
to keep adding energy into the system.

A damper, instead of storing energy dissipates energy. Since the damping force is proportional
to the velocity, the more the motion the more the damper dissipates the energy. Therefore a
point will come when the energy dissipated by the damper will equal the energy being fed in by
the force. At this point, the system has reached its maximum amplitude and will continue to
vibrate at this amplitude as long as the force applied stays the same. If no damping exists, there
is nothing to dissipate the energy and therefore theoretically the motion will continue to grow to
infinity.

Such catastrophic resonance can be witnessed frequently, in, for example, the failure of
complete aircraft wing structures during control surface “flutter”, failure of helicopter structural
components, and even the collapse of road bridges in gale force winds, as experienced at
Tacoma Bridge on November 7th, 1940.

Design Implications of Resonance


Designers of aircraft must be seriously concerned about the phenomenon of resonant frequency
because if a certain component of an aeroplane or helicopter is caused to vibrate at its resonant
frequency the amplitude of the vibration can become very large and the component will destroy
itself by vibration.

Module 2.2 Mechanics 2-95


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Let us examine the case of a helicopter which has a tail boom with a natural or resonant
frequency of 1 Hz. That is, if you were to strike the boom with your fist it would oscillate once
each second. The normal rotational speed of the rotor is 400 RPM and the helicopter has 3
blades on its main rotor. Each time a rotor blade moves over the tail boom the blade is going to
cause a downward pulse of air to strike the tail boom. The designer must determine the speed
at which the pulses will be equal to the resonant frequency of the boom. One cycle per second
is equivalent to 60 cycles/minute. Since each of the three blades causes a pulse each
revolution, there will be 3 x 60 or 180 pulses/minute. Therefore a rotor speed of 180 RPM would
be critical and the pilot would be warned against operating at that speed. Since the boom also
has a secondary, or overtone, resonant frequency of twice the fundamental resonant frequency,
360 RPM would also have to be avoided but would not be as critical as 180 RPM. The third
frequency of concern would be 3 x 180 or 540, but that is above the rotor operating speed, so is
not a problem.

The natural frequency of vibration is also an extremely important consideration in designing the
wings, horizontal and vertical stabilizers of an aircraft. The designer must be certain that the
resonant frequency when the surface is bent is different from that resonant frequency when it is
twisted. If that is not the case, an aerodynamic interaction with the elasticity of the surface can
result in “flutter” which can cause the surface to fracture in a fraction of a second after it begins.

Harmonics
The harmonic of an oscillation is a component frequency of the oscillation that is a multiple of its
natural frequency (known as the fundamental frequency). For example, if the fundamental
frequency is f, the harmonics have frequency 2f, 3f, 4f, etc. The harmonics have the property
that they are all periodic at the input frequency.

Thus, if an oscillating body (e.g. a spring/mass system) can be oscillated by an excitation input
of frequency equal to its natural frequency (the ‘fundamental frequency’), it will also be
oscillated at frequencies that are harmonics of that natural frequency.

2-96 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Problems

1. A pendulum has a length of 0.7m. What is its frequency of oscillation, and how long
will it take to oscillate 10 times?

2. A pendulum has a mass of 0.05 slugs. It takes 15 seconds to oscillate 10 times.


What is its length?

3. A mass of 0.4 kg oscillates freely on the end of a spring. The stiffness of the spring is
2 N/m. What is its natural frequency of oscillation and its time period?

4. A ball on the end of a spring bounces such that it nearly hits the floor 30 times in a
minute. The spring has a stiffness of 0.5 lb./in. What is the value of the mass of the
ball?

Module 2.2 Mechanics 2-97


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-98 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Answers

1. 0.6 Hz, 16.8 seconds

2. 0.625 m

3. 0.36 Hz, 2.8 seconds

4. 0.05 slugs

Module 2.2 Mechanics 2-99


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-100 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Simple Machines and the Principle of Work

The definition of work is as follows:

W = FD cos θ

The symbol for “distance” has been switched from S to D, to emphasize that we are dealing with
distances in our treatment of simple machines.

The angle (θ) in this definition is the angle between the direction of the force vector and the
direction of the displacement vector.

In this chapter, we will assume that in all the cases we will study the force and displacement
vectors act in the same direction. This implies that the angle (θ) is a 0° angle and since the
cosine of a 0o angle equals one, the equation for work becomes the simple equation:

W = FD

In this chapter, we will study six simple machines:

• The lever
• The pulley
• The wheel and axle
• The inclined plane
• The screw
• The hydraulic press

General Theory of All Machines


In discussing machines, we will assume that there is an object on which work is to be done. We
will call this object the load. In most cases, it is required that the load be raised a certain
distance in a gravitational field. For example, we wish to put cement blocks originally on the
ground into the bed of a truck.

A machine is a device for doing this work. The input work is, by definition, the work done by the
worker, that is, the force applied by the worker multiplied by the distance through which the
worker’s force acts. The output work is, by definition, the force that actually acts on the load
multiplied by the distance the load is raised.

We note that one way to do work is to do it directly. For example, it is possible for the worker to
raise each cement block directly to the truck bed. This is possible but can be difficult if each
block weighs, say, 175 lbs. In this case it would be better to use a machine since a machine
usually decreases the force supplied by the worker and increases the distance through which
his force acts.

Module 2.2 Mechanics 2-101


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
In the equations which follow, the subscript “o” will stand for output and the subscript “i” will
indicate input. We will use the following defining equations:

W o = FoDo

W i = F iD i

It is important to realize that there is no perfect machine. In our real world, on our earth, there is
always some friction. We always have, at least, air resistance. In addition, there is friction due to
the nooks and crannies that we would see if we inspected the surfaces of our machine parts
with a high-powered microscope.

Because of the constant presence of friction the input work is always greater than the output
work. Some of the input work is not useful work but serves to produce sound energy (a squeak),
light energy (a spark), or heat energy.

We will use the symbol “W f” to represent work lost because of friction.

Wi = W o + Wf

We define two kinds of “mechanical advantage”. The actual mechanical advantage (AMA) is
the ratio of the output force to the input force. This actual mechanical advantage tells us how
much easier it is for the worker. The ideal mechanical advantage (IMA) is the mechanical
advantage that would exist if there were no friction in the machine. It is the ratio of input
distance to the output distance.

Fo Di
AMA = IMA =
Fi Do

The ideal mechanical advantage of a machine can always be determined by measurements


made on the machine itself.

The efficiency (Eff) of a machine is the ratio of the output work to the input work.

Wo
Eff =
Wi

Fo D o
Eff =
FiD i

The efficiency can be expressed as a decimal or as a percentage. For example, if the efficiency
is calculated as 0.78, we can expressed it as 78%.

2-102 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
One final point should be made regarding efficiency. There is no machine that is 100% efficient.
We always have some friction. However, sometimes we assume that there is no friction and that
the machine is perfect or ideal! if a problem says that the efficiency is 100%, we are doing a
make-believe problem. This kind of a problem is not meaningless, however, because it tells us
the best that this machine can do. In this ideal case the AMA equals the IMA.

EXAMPLES:

A worker is able to raise a body weighing 300 lbs. by applying a force of 75 lbs. What is the
AMA of the machine that he is using?

Fo 300 lbs
AMA = = =4
Fi 75 lbs

A worker applied his force through a distance of 15 ft. The load is raised a distance of 2.5 ft.
What is the IMA of the machine that he used?

Di 15 ft
IMA = = =6
D o 2.5 ft

The actual mechanical advantage of a machine is 8 and the efficiency of this machine is 78%.
What is the ideal mechanical advantage?

AMA 8
IMA = = = 10.3
Eff 0.78

A worker uses a machine to raise a load of 500 lbs. a distance of 2 ft. He does this by applying
a force of 100 lbs. through a distance of 12 ft. What was the efficiency of the machine?

500 lbs
Method 1: AMA = =5
100 lbs

12 ft
IMA = =6
2 ft

AMA 5
Eff = = = 83%
IMA 6
Method 2:
W o = (500 lbs)(2 ft) = 1,000 ft.lbs.

W i = (100 lbs)(12 ft) = 1,200 ft.lbs.

W o 1,000 ft.lbs.
Eff = = = 0.83 = 83%
Wi 1,200 ft.lbs.

Module 2.2 Mechanics 2-103


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
We will next consider six simple machines. In each of these cases the IMA is expressed, not as
the ratio Di/Do, but in some other manner. We will study the geometry of each of these simple
machines to determine how to express the IMA in some simple equation.

The Lever
Consider the diagram in figure 2.42. Note that the
lever always pivots about some point called the
fulcrum. The input force (F1) is downward force and
in our diagram, is applied at the right end of the
lever. This input force gives rise to an upward force
at the left end in our diagram. This upward force
causes the load to be raised and is called “Fo”.

Figure 2.42: Simple lever


system

In figure 2.43. note that the input force acts through


a distance (Di) and the load is raised a distance
(Do).
Figure 2.43: Distances moved in a
simple lever system

The distance from the input end of the lever to the fulcrum is called the input lever arm (Li) and
the distance from the output end to the fulcrum is called the output lever arm (L0).

Recall that:

Di
IMA =
Do
However, figure 2.43 shows that the ratios of lever arms and distances are equal:

Di L
= i
Do Lo

Since it is much easier to measure lever arms that the distances of rotation, we always use the
ratio on the right hand side of the above equation to express the IMA of a lever.

Li
(Lever) IMA =
Lo

2-104 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
There are three classes of levers:

• 1st Class: The fulcrum is between the load and the applied force. Examples are the
claw hammer, scissors, and crowbar.
• 2nd Class: The load is between the fulcrum and the applied force. Examples are the
nutcracker and wheelbarrow.
• 3rd Class: The applied force is between the load and the fulcrum. An example is ice
tongs.

In a third class lever, the IMA is less than one. There is no force advantage. However, there is a
speed advantage. The work can be done in less time.

The Pulley
Some pulleys are firmly attached to an overhead support while other pulleys move up or down
with the load. We will refer to pulleys as “fixed” or “movable”.

In figure 2.44 (A), we have shown a single fixed pulley. If a length of pulley cord (Di) is pulled
down by a worker, the load will be raised a distance (Do). We see from the diagram that these
distances equal each other. Therefore we conclude that the IMA of this type of pulley is one. For
example, it would take 100 lbs. of force to raise a 100 lbs. load. The advantage of using this
type of pulley is that the worker is able to pull down on the pulley cord and in this way an
upward force is applied to the load. We say that a single fixed pulley is a “direction changer”.

Figure 2.44: Simple pulley systems


In figure 2.44 (B), there is a single movable pulley. A study of the diagram shows that Di is
always twice Do. For example, if the load is to be raised 2 ft. the worker must pull in 4 ft. of cord.

Note also that there are 2 strands supporting the load. The IMA of a single movable pulley is 2.

Module 2.2 Mechanics 2-105


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
In figure 2.44 (C), there is a single movable pulley and a single fixed pulley. The fixed pulley
again serves to change the direction of the input force. The IMA is still 2. Note also that there
are again 2 strands supporting the load.

We conclude that the IMA of a pulley equals the number of strands supporting the load.

(Pulley) IMA = the number of strands supporting the load

Several other examples of various types of pulley blocks are shown in figure 2.45.

Figure 2.45: More complex pulley systems

2-106 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
The Wheel and Axle

Note that one cord is wrapped around the axle of


radius (r). The load is attached to this cord. Another
cord is wrapped around the wheel of radius (A). The
worker applies his force to this second cord.

Both wheel and axle turn together. This means that if


the wheel rotates through one revolution the axle
also turns through one revolution.

Let us suppose that the worker pulls in a length of


cord equal to one circumference of the wheel (D1)
The load will be raised a distance equal one
circumference of the axle, (Do).

Di 2πR R
Figure 2.46: Wheel and IMA = = =
axle Do 2πr r

R
(Wheel and Axle) IMA =
r

The Inclinked Plane

In the inclined plane shown in figure 2.47 we note


that the worker slides the load up the incline. The
input distance (Di) is therefore equal to the length
of the incline (L). The effect of this is that the load
is raised a distance (h). This means that the
output distance (Do) equals h also.

Di L 1
IMA = = =
Figure 2.47: Inclined plane Do h h/L

We note that the sine of the angle of inclination (θ) is also h/L. Therefore, we can write the
expression for the IMA as follows:

1
(Inclined Plane) IMA =
sinθ

Module 2.2 Mechanics 2-107


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
The Screw Jack

The pitch of the screw (p) is the distance between adjacent


threads (see figure 2.48). As the handle is turned through
one revolution, a distance given by
2 π r ft., the load is raised a distance of one pitch.

Therefore, we have the relation:

2 π r
(Screw Jack) IMA =
p
Figure 2.48: The screw jack
A screw Jack has a great deal of friction. Therefore its efficiency is usually very low. However,
the distance through which the input force acts in comparison to the pitch is usually very large.
This gives a screw jack a large mechanical advantage.

The Hydraulic Press


A cross section of an hydraulic press is shown in figure 2.49. The small rectangles are cross
sections of the circular input and output pistons. Usually, we talk about the areas of the input
and output pistons (Ai and Ao). We note that the smaller of the two pistons is the input piston
(radius = r) and, of course, the larger piston is the output piston (radius = R).

An hydraulic press is filled with some fluid (gas or liquid). This fluid exerts a common fluid
pressure throughout the device.

Figure 2.49: The hydraulic press

As the smaller piston moves downward a distance (di) the larger piston moves upward a
distance (do). We recall that the volume of a cylindrical shape is equal to the area of the circular
base x the height. Also, a volume of fluid is transferred from the input (left) cylinder to the output
(right) cylinder. The volume of fluid is constant since the pressure is constant. Therefore, we
can write the equation:

π r 2di = π R 2do

2-108 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
We can cancel the common factor (it) and rearrange the equation. We obtain:

di R2
= 2
do r

The left member of this equation is, by definition, the IMA. Therefore, the IMA is also equal to
the right member of this equation. Thus, we can finally say that:

R2
(Hydraulic Press) IMA =
r2

We have obtained equations for the IMA of each of the six simple machines. We will do an
example of a typical problem dealing with machines. Note that any one of the six could be
chosen as an example. In the problems that follow the example, be sure to use the correct
formula for the IMA.

EXAMPLE:
The radius of the wheel in a windlass (wheel and axle) is 3.5 ft. and the radius of the axle is
0.27 ft. The efficiency of the machine is 60%. What load can be lifted by this machine by using
a force of 75 lbs.?
3.5ft
IMA = = 13.0
0.27ft

AMA = (Eff) (IMA)

AMA = (0.60) (13.0) = 7.8

Fo = (AMA) (Fi)

Fo = (7.8) (75 lbs.)

Fo = 585 lbs.

EXAMPLE:
An inclined plane has a 32° angle of incline. A for ce of 50 lbs. Is required to slide a 90 lbs.
load up the incline. What is the efficiency of this machine?
1
IMA = = 1.89
sin 32 o

90 lbs
AMA = = 1 .8
50 lbs

AMA 1.8
Eff = = = 0.95 = 95%
IMA 1.89

Module 2.2 Mechanics 2-109


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-110 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Problems

1. It takes a force of 80 lbs. to raise a body that weighs 240 lbs. What is the actual
mechanical advantage of the machine that was used?

2. A load is raised a distance of 6 ft. by a force acting through a distance of 18 ft. What is
the ideal mechanical advantage of the machine that was used?

3. What is the efficiency of a machine having an IMA of 7 and an AMA of 5?

4. A load weighing 120 lbs. is raised a distance of 4 ft. by a machine. The worker using the
machine exerts a force of 50 lbs. through a distance of 12 ft. What was the efficiency of
the machine?

5. The radius of the wheel of a windlass is 4.0 ft. and the radius of the axle is 0.2 ft. The
machine is 75% efficient. What force must be exerted to raise a load of 500 lbs. with this
machine?

6. The large piston of an hydraulic press has area 1.5 ft2. and the small piston has area
0.30 ft2. Assume that the machine is 100% efficient. What load can be raised by a force
of 75 lbs.?

7. A pulley system has four strands supporting the load. A force of 55 lbs. is needed to raise
a load of 200 lbs. What is the efficiency of this pulley system?

8. A light aircraft has a hydraulic braking system. Each rudder pedal is connected to a
master cylinder which provides braking for one of the main landing gear wheels. Each
master cylinder has a radius of 1/4-inch. The cylinder on the wheel has a radius of 1.0
inch. If the system is 95% efficient and the pilot exerts a force of 55 lbs. on the pedal,
how much force is exerted on the brake disc by the wheel cylinder?

Module 2.2 Mechanics 2-111


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-112 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Answers

1. 3

2. 3

3. 71%

4. 80%

5. 33.3 lbs.

6. 375 lbs.

7. 90.9%

8. 836 lbs.

Module 2.2 Mechanics 2-113


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-114 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Dynamics
Newton’s Laws
The rapid advance in aviation in the first half of the last century can be attributed in large part to
a science of motion which was presented to the world three centuries ago by Sir Isaac Newton,
a British physicist. Published in 1686, Newton’s treatise on motion, The Principia, showed how
all observed motions could be explained on the basis of three laws. The applications of these
laws have led to great technological advances in the aerodynamics, structure, and power plant
of aircraft. It is safe to say that any future improvements in the performance of aircraft will be
based on these laws. This chapter will be devoted to Newton’s laws, examining some of their
applications in aviation.

Newton’s First Law


The old magician’s trick of pulling a cloth out from under a full table setting is not only a
reflection of the magician’s skill but also an affirmation of a natural tendency which dishes and
silverware share with all matter. This natural tendency for objects at rest to remain at rest can
be attested to by any child who ever tried kicking a large rock out of his path. It is also a well
known fact that once a gun is fired, the command “stop” has no effect on the bullet. Only the
intervention of some object can stop or deflect it from its course. This characteristic of matter to
persist in its state of rest or continue in whatever state of motion it happens to be in is called
inertia. This property is the basis of a principle of motion which was first enunciated by Galileo in
the early part of the 17th century and later adopted by Newton as his first law of motion.

The first law of motion is called the law of inertia. It can be summarized:

A body at rest remains at rest and a body in motion continues to move at constant
velocity unless acted upon by an unbalanced external force.

The importance of the law of inertia is that it tells us what to expect in the absence of forces,
either rest (no motion) or straight line motion at constant speed. A passenger’s uncomfortable
experience of being thrown forward when an aircraft comes to a sudden stop at the terminal is
an example of this principle in action. A more violent example is the collision of a vehicle with a
stationary object. The vehicle is often brought to an abrupt stop.

Unconstrained passengers continue to move with the velocity they had just prior to the collision
only to be brought to rest (all too frequently with tragic consequences) by surfaces within the
vehicle (dashboards, windshields, etc.).

A less dramatic example of Newton’s first law comes from the invigorating activity of shovelling
snow. Scooping up a shovel full of snow, a person swings the shovel and then brings it to a
sudden stop. The snow having acquired the velocity of the shovel continues its motion leaving
the shovel and going off onto the snow pile.

Module 2.2 Mechanics 2-115


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Newton’s Second Law
A Learjet accelerates down the runway a distance of 3,000 feet, takes off and begins its climb at
6,000 feet per minute quickly reaching a cruising altitude of 35,000 feet, where it levels off at a
speed of 260 knots. Subsequently, the plane may have to perform a variety of manoeuvres
involving changes in heading, elevation, and speed. Every aspect of the aeroplane’s motion is
governed by the external forces acting on its wings, fuselage, control surfaces and power plant.
The skilled pilot using his controls continually adjusts these forces to make the plane perform as
desired.

The interplay between force and motion is the subject of Newton’s second law. An
understanding of this law not only provides insight into the flight of a plane, but allows us to
analyse the motion of any object.

While Newton’s first law tells us that uniform velocity is to be expected when an object moves in
the absence of external forces, the second law states that to have a change in speed or
direction an unbalanced force must act on the object. Using acceleration to describe the change
in motion of an object, the second law can he expressed

Fnet = ma

In words, the second law states that a net or unbalanced force acting on an object equals the
mass of the object times the acceleration of that object.

Here, the net force is the total force acting on the object, obtained by adding vectorially all of the
forces influencing the object. The mass is a scalar quantity. However, both the net force and the
acceleration are vector quantities. Mathematically, this means that they must always point in the
same direction. That is, at each instant the acceleration is in the same direction as the net force.

Before we consider cases where the net force acting on a body is not zero, it is most important
to understand that sometimes the net force acting on a body is zero. The vector sum of the
forces acting on the body in the x-direction is zero and the vector sum of the forces acting on
the body in the y-direction is also zero. In this case we say that the body is in equilibrium. From
the law, net force equals mass times acceleration, we know that since the net force is zero the
acceleration is also zero. Zero acceleration means that the velocity of the body in not changing
in direction or in magnitude. This means that the body is moving in a straight line with constant
speed or it has the constant speed, zero (it is at rest). If we observe that a body is at rest we
know that all of the forces on this body are balanced. Similarly, if a body is moving in a straight
line with constant speed, all of the forces acting on this body are balanced.

For example, if a plane is travelling on a straight stretch of runway at constant speed, there are
four forces acting on this plane: the earth is pulling down on the plane (its weight), the earth is
pushing up on the plane (the normal force), the engine is giving a forward thrust to the plane,
and frictional forces (air resistance, tires on runway. etc.) are acting backward. This is illustrated
in figure 2.50.

2-116 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Next, we must consider some examples where the net force acting on a body is not zero. The
body is accelerating. The body is experiencing a change in its direction or in its speed or both.
As a first example, a plane accelerating down a runway gets a change in velocity in the direction
of its motion. This is the same direction as the thrust provided by the power plant.

In figure 2.51, note that the thrust is greater than the frictional forces. The net forward force is
the thrust minus the friction. It is this net forward force that results in the acceleration of the
plane.

Figure 2.50: The four forces acting on an aeroplane

Figure 2.51: Thrust, drag and acceleration forces

Module 2.2 Mechanics 2-117


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Newton’s Third Law
Newton’s third law is sometimes referred to as the law of action and reaction. This law focuses
on the fact that forces, the pushes and pulls responsible for both the stability of structures as
well as the acceleration of an object, arise from the interaction of two objects. A push, for
example, must involve two objects, the object being pushed and the object doing the pushing.

Every action has an equal and opposite reaction

The third law states that no matter what the circumstance, when one object exerts a force an a
second object the second must exert an exactly equal and oppositely directed force on the first.
An apple hanging from a tree is pulled by the earth with a force which we call its weight.
Newton’s third law tells us that the apple must pull back on the earth with an exactly equal force.
The weight of the apple is a force on the apple by the earth, directed downward. The force
which the apple exerts back on the earth, is a pull on the earth directed upward. Another force
acting on the apple is the upward pull exerted by the branch. The law of action and reaction tells
us that the apple must be pulling down on the branch with the same magnitude of force.
People are often confused by this principle because it implies, for instance, that in a tug of war
the winning team pulls no harder than the losing team. Equally enigmatic is how a horse and
wagon manage to move forward if the wagon pulls back on the horse with the same force the
horse pulls forward on the wagon. We can understand the results of the tug of war by realizing
that the motion of the winning team (or losing team) is not determined exclusively by the pull of
the other team, but also the force which the ground generates on the team members feet when
they “dig in”. Recall, it is the net force, the sum of all of the acting forces which determines the
motion of an object.

The results of a “tug of war” can be quite different if the “winning team”, no matter how big and
strong, is standing on ice while the “losing team” is able to establish good solid footing on rough
terrain.

Similarly, the horse moves forward because the reaction force which the ground exerts in the
forward direction on its hooves is greater than the backward pull it receives from the wagon. By
focusing now on the wagon, we see that it moves forward because the forward pull of the horse
is greater than the backward pull of friction between its wheels and the ground.

Figure 2.52: Gravitational force between two objects

2-118 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Figure 2.53: Equal and opposite forces of Newton’s 3rd Law

One of the main difficulties people have with the third law comes from not realizing that the
action and reaction forces act on different objects and therefore can never cancel. Another
difficulty comes from forgetting that the motion of an object is determined by the sum of all of
the forces acting on that object.

In canoeing or rowing, a paddle is used to push water backward. The water reacts back on the
paddle generating a forward force which propels the boat.

Consider now a propeller as shown in figure 2.55.

The plane of rotation of the propeller is assumed to be perpendicular to the plane of the paper.
The flow of air is from left to right. We can imagine the action of the propeller is to take a mass
(m) of air on the left and accelerate it from some initial velocity (u) to a final velocity (v) to the
right of the propeller. The acceleration of this air mass requires a force which is provided by the
propeller. The air mass, in turn, reacts with an equal and

Figure 2.54: Equal and opposite forces of an oar

Figure 2.55: Action of a propeller

Module 2.2 Mechanics 2-119


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
opposite force on the propeller. This reaction force of the air on the propeller provides the thrust
for a propeller driven plane. The acceleration of the air mass is:

v −u
a=
t

Substituting this into Newton’s second law, we find for the net force on the air mass:

v −u
F=m
t

Both of the velocities (u and v) are the velocities relative to the plane of rotation of the propeller.
The time (t) is the time involved in accelerating the air mass from u to v.

By Newton’s third law, the thrust, which is the force the air mass exerts back on the propeller, is
equal in magnitude to F. Therefore, the thrust (T) is given by:

u−v
T=m
t

Recall that we have a symbol for “change in”, this means that we can write the above formula
as:

∆V
T=m
t

The velocities of the air mass are relative to the plane, and therefore change as the plane’s
speed changes. Also the time involved in accelerating the air mass changes with the speed of
the plane. This causes considerable variation in the thrust provided by a propeller.

EXAMPLE:

Each second a propeller accelerates an air mass of 12.2 slugs from rest to a velocity of 137
ft./sec. How much thrust is provided?

137 ft / sec − 0
T = (12 .2 slugs )
1 sec

T = 1,670 lbs.

2-120 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
In contrast to the reciprocating engine driven propeller which imparts a small change in velocity
to a relatively large mass of air, a turbojet induces a large change in velocity to a relatively small
mass of air. Here, the sole action of the Jet engine is considered to be the intake of a mass of
air at some velocity (u) and its exhaust at a higher velocity (v).

Figure 2.50 is a sketch of a turbojet engine. The velocity (u) in the figure denotes the relative
intake velocity and v denotes the exhaust gas velocity. The thrust formula which was obtained
above for the propeller will now be applied to a Jet engine. The thrust formula above can be
rewritten:

m m
T= v− u
t t

T = Gross thrust – Ram drag

The gross thrust is provided by the exhaust gases. The ram drag of the incoming air is due to
the speed of the aeroplane. The effect of the ram drag is to reduce the thrust provided by the
engine as the speed of the plane increases.

u v

Figure 2.57: Principle of the turbojet engine

EXAMPLE:

The Pratt & Whitney J60 has a mass air flow of 23 kg/sec. During a static test (initial velocity =
0) the exhaust velocity was measured to be 580 m/sec. Determine the thrust produced.

Note that the ram drag is zero since v1 is zero. Therefore, the thrust is equal simply to the
gross thrust.

m
T= v
t
23 kg
Substituting the given values we have: T= (580 m / sec) = 13,300 N
1 sec

Module 2.2 Mechanics 2-121


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
EXAMPLE:

What would the thrust have been if the J60 of the previous example had been in a plane
moving at 250 knots? Assume the same mass flow and exhaust velocity.
Note that the ram drag is not zero in this case. In order to calculate this ram drag we must use
the formula:
m
ram drag = u
t

Before substituting, we must express the initial velocity in m/sec.

1 .668 ft / sec 0 .30480 m


u = 250 knots × ×
1 knot 1 ft
= 129 m/sec.

23 kg
ram drag = 129 m / sec
1 sec

ram drag = 2,970 N

T = gross thrust - ram drag

T = 13,300 N - 2,970 N

T = 10,300N

EXAMPLE:

During a static test (initial velocity = zero), a Pratt & Whitney J75 produced a thrust of 16,000
lbs. with an air mass flow of 8.23 slugs/sec. Determine the exhaust gas velocity of the engine.

Since u is zero, the ram drag is zero and T = gross thrust.

m
T= v
t
We solve for the final velocity:
T
v=
m
t

16,000 lbs
v=
8.23 slugs / sec

v = 1,940 ft./sec.

2-122 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
The air intake velocity of a turbojet will be approximately equal to the airspeed of the plane. Let
us again examine the thrust formula.

m
T= (v − u)
t

It can be seen that the thrust may be increased in two ways, either by increasing the air mass
flow through the engine (M/t) or increasing the exhaust gas velocity (v).

EXAMPLE:

A French Dassault Falcon 30 is powered by two Lycoming ALF 502 turbofan engines. Flying
at sea level with a velocity of 154 m/sec. the air intake velocity is 154 m/sec. and the air
exhaust velocity is 224 m/sec. The airflow through each engine is 109 kg per second.
Determine the thrust of each engine.

109 kg
T= ( 224 m / sec − 154 m / sec)
1 sec

T = 7,630 N

EXAMPLE:

A Lockheed Jet Star is equipped with four Pratt & Whitney JT12 engines. Cruising at 220
knots, each engine was found to be providing 1,420 lbs. of thrust. If the airflow through each
engine was 1.55 slug/sec., what was the exhaust gas velocity?

U = 220 knots = 371 ft/sec.

Tt
= v −u
m

Tt
v= +u
m

(1,420 lbs)(1 sec)


v= + 371 ft / sec
1.55 slug

v = 1,290 ft./sec.

Module 2.2 Mechanics 2-123


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-124 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Motion in a Circle Centripetal
A ball whirled in a circle experiences an Force
acceleration toward the centre of the
circle.

This can be proven by considering that


the ball is continually changing direction
as it moves in a circle. Newton’s first
law tells us that the ball would prefer to
follow a straight path, and that for it to
deviate from a straight path, a force
must be applied to it.

It is a direct result of Newton’s first law


that a hammer thrower (Figure 2.58)
must continually pull towards the centre
of rotation, applying his full
weight to make the hammer accelerate
continually towards the centre of
rotation. As soon as the Figure 2.58: Centripetal force exerted by a
athlete stops applying the force towards hammer thrower
the centre (i.e. releases the hammer) the
hammer travels in a straight line, at a tangent to the circle.

This acceleration is in the same direction as the force which makes it move in a circle. This
force is called centripetal force (from the Latin meaning centre-seeking)

Since we have a constant change in the direction of the motion of the hammer, we have a
constant acceleration. This is called centripetal acceleration and can be calculated by the
square of the velocity divided by the radius of the circular path, thus:

V2
Centripetal acceleration =
R

Newton’s Second Law connects acceleration and force, by Force = Mass x Acceleration. Thus,
we can write the equation:

mV 2
Fnet = mac or Fnet =
R

Module 2.2 Mechanics 2-125


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Units of Force
The units which we will use in our discussion of Newton’s laws are the same as the units used
in the formula relating weight to mass (w = mg). These units are reviewed and summarized in
Table 2.4.

Each set of units, pound, slug, ft./sec.2) in the English system, or (Newton, kilogram, m/sec.2) in
the metric system is said to be consistent in the following sense: a force of 1 lb. when applied
to a mass of 1 slug gives it an acceleration of 1 ft./sec2.

Table 2.4: Units of Force, Mass and Acceleration

Similarly, a force of 1 Newton applied to a mass of 1 kilogram causes it to accelerate at 1


m/sec2.

Using Newton’s second law, we can write:

1 Newton = 1 kilogram m/sec2

and

1 pound = 1 slug ft./sec2

We note that Newton’s second law is correctly written as:

Fnet = ma

However, we often assume that the force acting on mass (m) is the net force. Thus, we usually
write the second law simply as:

F = ma

2-126 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
or, for circular motion,

mV 2
F= (Centripetal Force)
R

Newton’s second law when applied to bodies moving in a circular path states that the force
directed toward the centre of the path must equal the mass of the body times the square of the
speed of the body divided by the radius of the path. This force is called the centripetal (centre-
seeking) force.

EXAMPLE:

Find the acceleration of a 3 slug object acted upon by a net force of 1.5 lbs.

F
a=
m

1.5 lbs
a=
3 slugs

a = 0.5 ft./sec2
EXAMPLE:

A mass of 6 kilograms accelerates at 5 rn/sec.2 Find the force which is acting on this object.
F = ma

F = (6 kg) (5 m/sec2) = 30 N

Module 2.2 Mechanics 2-127


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-128 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Problems

1. Find the mass of an object which accelerates at 5 m/sec2 when acted on by a net force of
one Newton.

2. Find the acceleration of a 3 slug object experiencing a net force of 12 lbs.

3. Find the net force on a 5 slug object which is accelerating at 3ft/sec2

4. A Learjet Model 24 of mass 6,000 kg is observed to accelerate at the start of its takeoff at
4 m/sec2 . What is the net forward force acting on the plane at this time?

5. During a static test, a Continental engine driving a two blade constant speed propeller
was found to accelerate each second a mass of 140 kg from rest to a velocity of 40
m/sec. Determine the thrust on the propeller.

6. A Piper Archer ii has an Avco Lycoming engine driving a two blade propeller. Each
second 8 slugs of air are given a change in velocity of 122 ft/sec. How much thrust is
generated on the propeller?

7. The Garrett TFE 731 turbofan engine which powers the Rockwell Saberliner 65 under
static testing has an exhaust gas velocity of 321 m/sec and an airflow of
50 kg/sec. Find the static thrust of the engine.

8. A plane weighs 36,000 lbs. The forward thrust on the plane is 20,000 lbs. and the
frictional forces (drag) add up to 2,000 lbs. What is the acceleration of this plane? Hint:
Be sure to find the mass of the plane from its weight.

9. What centripetal force is needed to keep a 3 slug ball moving in a circular path of radius
2 feet and speed 4ft/sec.?

10. A boy is swinging a stone at the end of a string. The stone is moving in a circular path.
The speed of the stone is 5 ft/sec. and the radius of the path is 1.5 ft. What is the
centripetal acceleration of the stone?

Module 2.2 Mechanics 2-129


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-130 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Answers

1. 0.2 kg

2. 4 ft./sec2

3. 15 lb.

4. 24,000 N

5. 5600 N

6. 976 lb.

7. 16,050 N

8. 16 ft/sec2

9. 24 lb.

10. 16.67 ft/sec2

Module 2.2 Mechanics 2-131


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-132 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Friction
When a body rests on a horizontal surface or is dragged or rolled on such a surface there is
always contact between the lower body surface and the horizontal surface. This contact results
in friction. Friction is work done as the surfaces rub against each other. This work heats the
surfaces and always results in wasted work.

We need to define a force known as the normal force. A body resting on a horizontal surface
experiences two forces, the downward force due to the gravitational pull of the earth on this
body (the weight of the body), and the upward push of the surface itself on the body (the normal
force).

The weight (w) and the normal force (N) are equal to
each other.

There are three kinds of friction:


1. Static friction
2. Sliding friction
3. Rolling friction
Figure 2.58: Weight and its
reaction
Static Friction
Static friction (or ‘starting’ friction) is the force between two objects that are not moving relative
to each other. For example, static friction can prevent an object from sliding down a sloped
surface. The coefficient of static friction, typically denoted as µs, is usually higher than the
coefficient of kinetic friction. The initial force to get an object moving is often dominated by static
friction.

Another important example of static friction is the force that prevents a car wheel from slipping
as it rolls on the ground. Even though the wheel is in motion, the patch of the tire in contact with
the ground is stationary relative to the ground, so it is static rather than kinetic friction.

The maximum value of static friction, when motion is impending, is sometimes referred to as
limiting friction, although this term is not used universally

Rolling Friction
Rolling friction is the frictional force associated with the rotational movement of a wheel or other
circular objects along a surface. Generally the frictional force of rolling friction is less than that
associated with kinetic friction. One of the most common examples of rolling friction is the
movement of motor vehicle tyres on a road, a process which generates heat and sound as by-
products.

Kinetic Friction
Kinetic (or dynamic) friction occurs when two objects are moving relative to each other and rub
together (like a sled on the ground). The coefficient of kinetic friction is typically denoted as µk,
and is usually less than the coefficient of static friction. Since friction is exerted in a direction
that opposes movement, kinetic friction usually does negative work, typically slowing something
down. There are exceptions, however, if the surface itself is under acceleration. One can see

Module 2.2 Mechanics 2-133


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
this by placing a heavy box on a rug, then pulling on the rug quickly. In this case, the box slides
backwards relative to the rug, but moves forward relative to the floor. Thus, the kinetic friction
between the box and rug accelerates the box in the same direction that the box moves, doing
positive work.

Examples of kinetic friction:

• Sliding friction is when two objects are rubbing against each other. Putting a book flat
on a desk and moving it around is an example of sliding friction
• Fluid friction is the friction between a solid object as it moves through a liquid or a gas.
The drag of air on an aeroplane or of water on a swimmer are two examples of fluid
friction.

Calculating Friction
In all cases, the friction equation is the same.

F = µN

The symbol “µ” (the Greek letter mu) is called the coefficient of friction.
Every pair of flat surfaces has two different coefficients of friction:

Coefficients of Friction
Material µstart µslide
Steel on Steel 0.15 0.09
Steel on ice 0.03 0.01
Leather on Wood 0.5 0.4
Oak on Oak 0.5 0.3
Rubber on Dry Concrete 1.0 0.7
Rubber on Wet Concrete 0.7 0.5
Table 2.5: Some examples of Coefficients of Friction

The coefficient of starting friction — µstart


The coefficient of sliding friction — µslide

Some values for the coefficients of starting and sliding friction are given in Table 2.5.

We note that the coefficients of sliding friction are less than the coefficients of starting friction.
This means that the force needed to start a body sliding is greater than the force needed to
keep a body sliding with constant speed.

When we deal with a body that rolls over a flat surface, we have another coefficient of friction to
consider: the coefficient of rolling friction.

2-134 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
The coefficients of rolling friction (µroll) are very small. Therefore, rolling friction is much smaller
than either starting or sliding friction. Some values are:

Rubber tires on dry concrete 0.02


Roller bearings 0.001 to 0.003

EXAMPLE:

A steel body weighing 100 lbs. is resting on a horizontal steel surface. How many pounds of
force are necessary to start the body sliding? What force is necessary to keep this body
sliding at constant speed?

W = N = 100 lbs.

F = µN

Force to start sliding motion = (0.15)(100 lbs.) = 15 lbs.

Force to keep body sliding = (0.09) (100 lbs.) = 9 lbs.

Module 2.2 Mechanics 2-135


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-136 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Problems

1. An aircraft with a weight of 85,000 lbs. is towed over a concrete surface. What force must
the towing vehicle exert to keep the aircraft rolling?

2. It is necessary to slide a 200 lb. refrigerator with rubber feet over a wet concrete surface.
What force is necessary to start the motion? What force is necessary to keep the motion
going?

Module 2.2 Mechanics 2-137


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-138 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Answers

1. 1,700 lbs.

2. 140 lbs. 100 lbs.

Module 2.2 Mechanics 2-139


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-140 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Work, Energy and Power

Work
Work is done on a body when a force acts through a distance. The definition of work involves
the force acting on the body (F) the distance through which this force acts (S) and the angle (θ)
between the force vector and the distance vector. The definition of work is:

W = FS cos θ

Very often the force vector and the distance vector act in the same direction. In this case, the
angle (θ) is a zero degree angle. If you check on your calculator, you will find that the cosine of
a zero degree angle is equal to one. This simplifies things in this case because then work is
simply equal to the product of force times distance.

The unit of work in the English system is the foot-lb. Note that the two units are multiplied by
each other. Students tend to write ft./lb. This is incorrect. The unit is not feet divided by pounds.
In the metric system, the unit is the Newton-meter or the Joule (J). Note that the Newton-meter
has a name, the Joule. The foot-lb. has no special name.

EXAMPLE:
A puck lies on a horizontal air table. The air table reduces the friction between the puck and
the table to almost zero since the puck rides on a film of air. A player exerts a force of 70 lbs.
on this puck through a distance of 0.5 feet, and he is careful that his force is in the same
direction as the distance through the force is applied. The player has done 35 ft.-lbs. of work
on the puck.

EXAMPLE:
A book weighing 8 pounds is raised a vertical distance by a student demonstrating work. The
book is raised 2 feet. The student has done 16 ft.lbs. of work.

EXAMPLE:
A sled is dragged over a horizontal snowy surface by means of a rope attached to the front of
the sled. The rope makes an angle of 28° with the h orizontal. The sled is displaced a distance
a 50 ft. The worker exerts a force of 35 pounds. How much work does the worker do? We
use the formula:

W = FS cos θ

W = (35 lbs.) (50 ft.) cos 28°

W = 1550 ft.lbs.

Module 2.2 Mechanics 2-141


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Sometimes the force and the displacement are in the opposite directions. This situation gives
rise to negative work. Note, in this case, the angle between the force and the displacement is a
180° angle. The cosine of 180° is negative one.

One example of negative work occurs when a body is lowered in a gravitational field. If a
student carefully lowers a book weighing 15 pounds through a distance of 2 feet, we note that
the displacement vector points downward and the force vector point upwards

W = FS cos θ

W = (15 lbs.)(2 ft.) cos 180o

W = (15 lbs.)(2 ft.)(-1)

W = -30 ft.lbs.

Energy
The concept of energy is one of the most important concepts in all of physical science. We often
hear of energy sources, alternate energy, shortage of energy, conservation of energy, light
energy, heat energy, electrical energy, sound energy, etc. What is the exact meaning of this
word, energy?

Sometimes energy is defined as the “capacity to do work”. This definition is only a partial
definition. However, it has the advantage of immediately relating the concept of energy to the
concept of work. These two ideas are intimately related to each other.

Energy is a quality that a body has after work has been done on this body. Once work has been
done on a body of mass (m) this body has energy. The body can then do work on other bodies.
Consider the following situation:

A body of mass (m) was resting on a horizontal air table. A player exerted a horizontal force (F)
on this mass through a distance (s). Since the angle between the force and the displacement
was a zero degree angle, the work done on this body was simply Fs. At the instant the player
removed his hand from the body we note two facts. The body accelerated while the force (F)
was acting on the body and the body has acquired a velocity (v) during this time of acceleration
(a). The body has moved through a distance (s) in time (t).

s = ½ at 2

Also note that the force (F) is related to the acceleration by the relation:

F = ma

We now look again at this body at the instant the force (F) has ceased acting. We note that
work (W) has been done on this body and that the body moves with speed (v).

W = Fs = (ma) ( ½ at 2)

2-142 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
W = ½ m(at) 2

Now we note that the speed obtained by the body during the time of acceleration is given by the
equation:

V = at

Therefore, we can substitute v for at in the equation above.

W = ½ mV 2

The equation we have obtained is the defining equation for a quantity known as kinetic energy.

Usually, we use the symbol “KE” for kinetic energy.

KE = ½ mV 2

After the work has been done on the mass (m) it moves off on the frictionless air table with this
kinetic energy. This body now is capable of doing work on other bodies that it contacts. For
example, it probably will strike the edge of the table. When this happens this kinetic energy will
be changed into other types of energy such as sound energy or heat energy.

We note that the initial kinetic energy of the mass (m) was zero. This is true because the body
was initially as rest. We can say that the work done on the body is equal to the change in the
energy of the body.

Gravitational Potential Energy


Another equally important situation where an agent easily can do work on a body (and thus give
the body energy) occurs when the agent raises a body vertically in a gravitational field, at the
surface of the earth.

In this case, the work done on the body again equals the force applied multiplied by the
distance the body is raised.

W = Fs

W = (weight of body) (distance raised)

We recall that w = mg. Also since the distance is a vertical distance we use the symbol “h” for
height. In our discussion we will assume that the symbol “h” always represents the vertical
distance of the body above the surface of the earth.

Therefore, we write:

W = mgh

Module 2.2 Mechanics 2-143


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Again we have a case where an agent did work on a body and the body has acquired “energy”.
This type of energy is known as gravitational potential energy. however, we usually symbolize it
as “PE”.

PE = mgh

If we neglect air resistance (which results in loss of energy to heat), we note that there is a
conservation of kinetic and potential energy of a body moving in a gravitational field. As a body
falls from a height (h) and moves closer to the surface of the earth, its potential energy
decreases and its kinetic energy in creases while it is falling. Therefore, there is an easy way of
finding the speed of a falling body during any instant of its fall.

The units for energy are the same as the units for work, the Joule (J) in the metric system and
the foot-pound in the English system.

EXAMPLE:
A body of mass 4 slugs is held by an agent at a distance of 6 ft. above the surface of the
earth. The agent drops the body. What is the speed of the body when it is on the way down
and at a distance of 2 feet above the earth’s surface?

We note that the initial potential energy is equal to the sum of the kinetic and potential
energies on the way down (wd).

PE = PEwd + KEwd

(4 slug) (32 ft/sec2) (6 ft.) = (4 slug) (32 ft/sec2) (2 ft.) + ½ (4 slug) v 2

EXAMPLE:
A body of mass, 10 kg, falls to the earth from a height of 300 m above the surface of the
earth. What is the speed of this body just before it touches ground?
PEi = KEf

(10 kg) (9.8 m/sec2) (300 m) = (10 kg) V2

2,940 m2/sec2 = ½ V 2

5,880 m2/sec2 = V2

V = 76.7 m/sec

The kinetic energy that the body has just before it reaches the ground immediately changes to
sound energy and heat energy on impact. It may also “squash” any body in its path or make an
indentation in the earth – this is strain energy (energy to deform).

2-144 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Power
Power is the rate of doing work. The more rapidly a piece of work can be done by a person or a
machine the greater is the power of that person or machine.

We define power by the following equation:

work force × dis tan ce


Power = =
time time

In symbols:

Fs
P=
t

In the English system the unit of power is the horsepower and in the metric system the unit is
the Watt.

Conversion factors exist giving information regarding these units.

ft.lbs. ft.lbs.
1 Horsepower = 550 = 33,000
sec . min .

1,000 Watts = 1 kilowatt

EXAMPLE:

An aircraft engine weighing 4,000 lbs. is hoisted a vertical distance of 9 feet to install it in an
aircraft. The time taken for this piece of work was 5 minutes. What power was necessary?
Give the answer in ft.lb./sec. and in horsepower.

Fs ( 4,000 lbs )( 9 ft )
P= =
t 300 sec s.

1 HP
P = 120 ft.lbs. / sec . ×
550 ft.lbs. / sec .

P = 0.218 HP

Module 2.2 Mechanics 2-145


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
EXAMPLE:

An elevator cab weighs 6,000 N. It is lifted by a 5 kW motor. What time is needed for the cab
to ascend a distance of 40 m?

Fs
t=
P

(6,000 N)( 40 m)
t=
5,000 Watts

t = 48 seconds

Alternate Form for Power


s
We can put our formula for power in another form by recognizing that is speed (v). This leads
t
to the formula:

Fs s
P= =F
t t

P = Fv

This form is particularly useful for obtaining an expression for the power output of a turbine
engine. These engines are ordinarily rated in terms of the thrust which they produce. To obtain
an expression for their power output it is necessary to multiply their thrust by the speed of the
plane. This thrust power, which is usually expressed in units of horsepower (THP, thrust
horsepower), can be obtained by multiplying the thrust in pounds by the speed in ft./sec. and
dividing by 550 where the conversion
1 HP = 550 ft-lbs./sec. is used. Thus:

thrust in lbs. × aircraft speed in ft. / sec .


THP =
550

Alternatively, we can take the speed of the aircraft in MPH and use the conversion
1 HP = 375 mi.lbs./hr. to obtain:

thrust in lbs. × aircraft speed in MPH.


THP =
375

2-146 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
EXAMPLE:

A gas turbine engine is producing 5,500 lbs. of thrust while the plane in which the engine is in-
stalled is travelling 450 MPH. Determine the THP.

(550 lbs )( 450 MPH )


THP = = 6,600 HP
375

It is important to note that while the thrust of a gas turbine engine may not vary much over a
particular range of aircraft speeds, the power must be recalculated each time the plane changes
its speed.

Module 2.2 Mechanics 2-147


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-148 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Problems

1. How much work is done by a person in raising a 45 lb. bucket of water from the bottom of
a well that is 75 ft. deep? Assume the speed of the bucket as it is lifted is constant.

2. A tugboat exerts a constant force of 5,000 N on a ship moving at constant speed through
a harbour. How much work does the tugboat do on the ship in a distance of 3 km?

3. A father has his 45 lb. son on his shoulders. He lowers the child slowly to the ground, a
distance of 6ft. How much work does the father do?

4. A 6 slug body has a speed of 40 ft/sec. What is its kinetic energy? If its speed is doubled,
what is its kinetic energy?

5. A 2 kg ball hangs at the end of a string 1 m in length from the ceiling of a ground level
room. The height of the room is 3 m. What is the potential energy of the ball?

6. A body of mass 3 slug is a distance of 77 ft. above the earth’s surface and is held there
by an agent. The agent drops the body. What is the speed of the body just before it hits
ground?

7. An aircraft of mass 4 tonnes lands at 30 m/s and the pilot immediately applies the brakes
hard. The brakes apply a retarding force of 2000N. How far will the aircraft travel before
it comes to rest.

8. A pile driver of mass 1000 kg, hits a post 3 m below it. It moves the post 10 mm. What
is the kinetic energy of the pile driver?

9. A pile driver of mass 1000 kg, hits a post 3 m below it. It moves the post 10 mm. With
what force does it hit the post when it hits the post?

10. An aircraft engine weighing 12,000 N is lifted by a 3.6 kW motor a distance of 10m. What
time was needed?

11. A hand-powered hoist is used to lift an aircraft engine weighing 3,000 lbs. a vertical
distance of 8 ft. If the worker required 4 minutes to do this job, what horsepower was
developed by the mechanic?

12. How long does it take a 5 kW motor to raise a load weighing 6,000 lbs. a vertical distance
of 20 ft.? (Hint: convert KW to ft.lb./sec first)

Module 2.2 Mechanics 2-149


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-150 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Answers

1. 3370 ft.lb.

2. 15,000,000 J

3. -270 ft.lb. (Note the negative sign!)

4. 4800 ft.lb; 19,200 ft.lb.

5. 39.2 J (or 40 of g = 10m/s2)

6. 70 ft./sec.

7. 900 m

8. 30,000 J

9. 3 MN

10. 33 sec.

11. 100 ft.lb./sec. 0.182 HP

12. 32.5 sec.

Module 2.2 Mechanics 2-151


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-152 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Momentum

Definition of Momentum
Momentum is a vector quantity defined as the product of mass times velocity. Note that velocity
(V) is also a vector quantity. We write the defining equation as:

Momentum = mV

Momentum is a very important quantity when we are dealing with collisions, because it is con-
served in all such cases.

Conservation of Momentum
In a collision, there are always at least two bodies that collide. We will deal only with collisions
of two bodies. We will also limit our discussion to collisions occurring in one dimension. Such
collisions are called “head-on” collisions.

At this time, we need to recall two of Newton’s laws. We need Newton’s second law,
F = ma, and Newton’s third law, which tells us that if two bodies collide, the force that the first
body exerts on the second body is equal in magnitude and opposite in direction to the force that
the second body exerts on the first body. Also recall that the acceleration (a) equals the change
in the velocity (symbolized by the Greek letter Delta, ∆) divided by the time (t).

Now let us visualize two bodies of masses, m1 and m2 on a one dimensional track.

If these two bodies collide, we have four different velocities to consider. We will name these
velocities very carefully.

V1’ = the velocity of body one before the collision


V1”= the velocity of body one after the collision

V2’ = the velocity of body two before the collision


V2” = the velocity of body two after the collision

From Newton’s two laws, we can conclude that:

∆V1 ∆V2
m1 = −m 2
t t
After cancelling ‘t’, we obtain:

m1(V1” - V1’) = - m2(V2” - V2’)

If we remove the parentheses, transpose terms, and switch left and right parts we obtain:

m1V1’ + m2V2’ = m1V1” + m2V2”

The equation tells us that the total momentum before the collision is equal to the total
momentum after the collision. Sometimes we say simply that “momentum is conserved”.

Module 2.2 Mechanics 2-153


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Recoil Problems
The simplest example of the conservation of momentum is in recoil problems.

EXAMPLE:

A boy and a man are both on ice skates on a pond. The mass of the boy is 20 kg and the
mass of the man is 80 kg. They push on each other and move in opposite directions. If the
recoil velocity of the boy is 80 m/sec., what is the recoil velocity of the man?

First we note that both the man and the boy are at rest before the collision occurs.

m1V1’ + m2V2’ = m1V1” + m2V2”

(20) (0) + (80) (0) = (20) (80) + (80)V2”

0 = 1,600 + 80V2”

-1,600 = 80V2”

V2” = - 20 m/sec.

The negative sign indicates that the man recoils in the opposite direction from the boy.

Collision Problems
Whenever two bodies collide, momentum is always conserved. This is simply the result of
applying Newton’s second and third laws as we have done in the preceding discussion.

Sometimes kinetic energy is also conserved in a collision. This happens when the bodies are so
hard that there is very little deformation of the bodies in the actual collision process. Billiard balls
are a good example. These collisions are known as elastic collisions. We will derive a formula
for determining the velocities of the bodies after the collision has occurred.

Another type of collision that we will discuss is the perfectly inelastic collision. In this type of
collision, the bodies are deformed so much that they actually stick together after the collision.
An example would be the collision of two masses of putty. We will also do some problems for
this type of collision.

2-154 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Inelastic Collisions
We use the conservation of momentum for dealing with this type of collision. As we have said,
the colliding bodies stick together after impact. Therefore, the equation is simply:

m1V1’ + m2V2’ = (m1 + m2) V”

Note that we use the symbol V” for the common velocity of the two bodies (which are now one
body) after the collision.

It is important to include the signs of the velocities of the bodies in setting up momentum
equations. As usual, we use a positive sign for east and a negative sign for west, a positive sign
for north and a negative sign for south.

EXAMPLE:

A truck of mass 1,550 kg is moving east at 60 m/sec. A car of mass 1,250 kg is travelling west
at 90 m/sec. The vehicles collide and stick together after impact.

What is the velocity of the combined mass after the collision has occurred?

V1’ = 60 m/sec.
m1 = 1,550 kg

V2’ = -90 m/sec.


m2 = 1,250 kg

m1V1’ + m2V2’ = (m1 + m2) V”

We will not include units in our substitution. However, we will note that the velocity, when we
obtain it, will be in m/sec.

(1,550) (60) + (1,250) (-90) = (1,550 + 1,250)V”

-19,500 = 2,800 V”

V” = -6.96 m/sec.

Since the calculated velocity has a negative sign, we conclude that the combined mass is
travelling west after the impact has occurred.

Our answer is that the wreckage starts moving west with a speed of 6.96 m/sec.
Sometimes the principle of conservation of momentum in the case of an inelastic collision can
be used by the police to determine the speed of a vehicle engaged in a head-on collision.

Module 2.2 Mechanics 2-155


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Sometimes the principle of conservation of momentum in the case of inelastic collision can be
used by the police to determine the speed of a vehicle engaged in a head –on collision.

Suppose that a large truck with a weight of 12,000 lbs. (mass of 375 slugs) travelling east with
an unknown velocity enters into a head-on collision with a smaller truck of weight 6.400 lbs.
(mass of 200 slugs) initially travelling west with a speed of 30 MPH (44 ft./sec.). The trucks stick
together in the collision and marks on the highway indicate that the wreckage travelled a
distance of 120 feet east. The condition of the roadway (amount of friction) indicates that the
wreckage would travel for a time of 4 sec. Determine the initial speed of the large truck.

The equation:

u+v
s= t
2

can be used to determine the initial velocity of the wreckage. Note that the final velocity of the
wreckage is zero.

2s 2(120 ft )
u= = = 60 ft. / sec .
t 4 sec

Next, we can use the conservation of momentum equation to determine the velocity of the large
truck at the instant of the impact. We will use the symbol V to represent this velocity.

(375 slugs) (V) + (200 slugs) (- 44 ft./sec.) = (575 slugs) (60 ft./sec.)

375V = 43,300

V = 115 ft./sec.

V = 78.4 MPH

Elastic Collisions
Elastic collisions are collisions that occur between bodies that deform very little in the collision.
Therefore we assume that no energy is lost. An example of such a collision is the collision
between pool balls.

In elastic collisions, both kinetic energy and momentum are conserved. In an ordinary elastic
collision problem, we know the masses and the velocities of two bodies that will collide. We
want to predict, by a mathematical calculation, the velocities the bodies will have after the
collision has occurred, the two unknowns. If we write the two conservation equations, we have
two equations in these two unknowns. It is possible to solve these two equations for these two
unknowns. However, one of the conservation equations, the energy equation, is a “second
order” equation. A “second order” equation contains the squares of the unknowns. This makes
the solution more difficult. Instead, we will use an algebraic trick! The two conservation
equations can be solved together producing a third equation. This third equation and the
momentum conservation equation provide the two first order equations that we will use in
solving elastic collision problems.

2-156 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
We will write the two conservation equations:

Conservation of Energy

(1) ½ m 1V1’2 + ½ m 2V2’2 = ½ m 1V1”2 + ½ m 2V2”2

Conservation of Momentum

(2) m1V1’ + m2V2’ = m1V1” + m2V2”

Divide (1) by ½:

(3) m1V1’2 + m2V2’2 = m1V1”2 + m2V2”2

Now in both (2) and (3), we will transpose some terms:

(4) m1V1’ - m1V1” = m2V2” - m2V2’

(5) m1V1’ - m1V1” = + m2V2” - m2V2’

Factorise (4) and (5):

(6) m1(V1’ - V1”) = m2(V2” - V2’)

(7) m1(V1’2 - V1”2) = m2(V2”2 - V2’2)

In (7), Factor again:

(8) m1(V1’ - V1”) (V1’ + V1”) = m2(V2” - V2’) (V2” + V2’)

Divide (8) by (6):

m1( V1 '− V1 " ) ( V1 '+ V1 " ) m 2 ( V2 "− V2 ' ) ( V2 "+ V2 ' )


=
m1 ( V1 '− V1 " ) m 2 ( V2 "− V2 ' )

After cancelling common factors, we obtain:

V1’ + V1” = V2” + V2’

Again we transpose terms:

Module 2.2 Mechanics 2-157


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
(9) V1’ - V2’ = V2” - V1”

In words, this equation says that the relative velocity of the balls before the collision is equal to
the relative velocity of the balls after the collision.

Equation (9) has been obtained algebraically from two equations, the conservation of
momentum and the conservation of energy. We use equations (2), the conservation of
momentum equation, and equation (9), called the relative velocity equation, to solve for the
velocities of the two bodies after an elastic collision.

We will rewrite these two important equations for future reference:

(2) m1V1’ + m2V2’ = m1V1” + m2V2”

(9) V1’ - V2’ = V2” - V1”

In using these two equations, the two unknowns are usually V1” and V2”, the velocities of the
two bodies after the collision has occurred. The known quantities are usually the two masses
and the velocities of the bodies before the collision. Also be careful to include the signs of the
velocities. If you forget to do this, you will always end up with incorrect results.

EXAMPLE:

A billiard ball of mass 2 kg is moving east at 3 m/sec. and undergoes an elastic collision
with another billiard ball of mass 3 kg moving west at 4 m/sec. Find the velocities of the
two balls after the collision.
m1 = 2 V1’ = 3 (east)
m2 = 3 V2’ = -4 (west)

Substitute in equation (2):

(2)(3) + (3)(-4) = 2V1” + 3V2”

(10) -6 = 2V1” + 3V2”

Substitute inequation (9):


3 - (-4) = V2” - V1”

(11) 7 = V2” - V1”

Rewrite equations (10) and (11) putting the unknowns in the left members and in order.

(10) 2V1” + 3V2”= -6


(11) -V1” + V2” = 7

We now have two equations and two unknowns. There are several methods of solving
such a system of equations. We will use the method of addition. In this method we multiply

2-158 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
either or both of the equations by constants to make the coefficient of one of the unknowns
in one of the equations a positive number and to make the coefficient of this same
unknown in the other equation a negative number of the same magnitude. We then add
the two equations to eliminate one of the unknowns. We then solve for the other unknown
by substituting in either equation.

We will multiply (11) by the number 2.

(12) -2V1” + 2V2” = 14

Add (10) and (12):

5V2” = 8

V2” = 1.6 m/sec.

Substitute this value back into (11):

-V1” + 1.6 = 7

V1” = -5.4 m/sec.

We note that we interpret a positive sign for the velocity as motion east and a negative sign
as motion west.

Our final result is that the 2 kg ball is moving west with a speed of 5.4 m/sec after the
collision and the 3 kg ball is moving east with a speed of 1.6 m/sec. after the collision.

Module 2.2 Mechanics 2-159


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-160 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Problems

1. A gun of mass 5 kg fires a bullet of mass 20 grams. The velocity of the bullet after firing,
is 750 m/sec. What is the recoil velocity of the gun?

2. An astronaut on a space walk has a mass of 5 slugs and is at rest relative to the space
station. She is working with a tool having a mass of 0.5 slug. She accidentally throws this
tool away from herself with a speed of 6 ft/sec. With what speed does the astronaut
recoil?

3. An automobile having mass 1,500kg is travelling east on an expressway at 30m/sec. It


overtakes a truck of mass 2,000kg also travelling east and moving with a speed of 25
m/sec. The automobile rear-ends the truck. The vehicles become locked together in this
collision and continue east. What is the velocity of this combined mass?

4. Two balls of putty become one mass of putty in a collision. The first, of mass 6 kg, was
originally moving east at 10 m/sec., and the second, of mass 4 kg was originally moving
west at 9 m/sec. What is the velocity of the total mass after the collision has occurred?

5. Due to a controller’s error two aircraft are directed to land in opposite directions on the
same runway in a fog. A Cessna 150 of mass 50 slugs and a Beechcraft Bonanza of
mass 80 slugs undergo a direct head-on collision. The Beech-craft Bonanza was
originally travelling north at a speed of 30 MPH and the Cessna was travelling south. The
wreckage travels a distance of 20 ft. south during a time of 3.6 sec. What was the original
speed of the Cessna?

6. A 3 kg ball is moving right with a speed of 3 m/sec. before a collision with a


2 kg ball originally moving left at 2 m/sec. What are the directions and speeds of the two
balls after the collision?

7. A 2 kg ball moving right at 5 m/sec. overtakes and impacts a 1 kg ball also moving right
at 2 m/sec. What are the speeds and directions of the two balls after the impact?

Module 2.2 Mechanics 2-161


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-162 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Answers

1. 3 m/sec.

2. 0.6 ft./sec.

3. 27 m/sec. East

4. 2.4 m/sec. East

5. 67.7 MPH
6. The 3 kg ball is moving left at 1 m/sec. and the 2 kg ball is moving right at
4 m/sec.

7. The 2 kg ball is moving right at 3 m/sec. and the 1kg ball is moving right at
6 m/sec.

Module 2.2 Mechanics 2-163


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-164 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Torque
Consider the diagrams 2.59 shown below. We define torque as the force (F) applied to a body
that is pivoted at a point (0) multiplied by the distance from the pivot point to the place where the
force is applied and multiplied by the sin of the angle between r and F. We will use the Greek
letter Tau (τ) for torque. The distance mentioned in the preceding sentence is called the lever
arm and symbolized by the letter r.

The defining equation is:

τ = r F sin θ

In the diagram , we note that θ = 90°. This is by far the most common case. Since
sin 90° = 1, this common case reduces to the more s imple equation:

τ= rF

However, it must be remembered that in those cases where θ is not 90°, the full equation must
be used. Note also that the unit for torque is the lb.ft. , lb.in. or the Nm.

Figure 2.59: Force acting at a


distance creates torque

Extensions
Figure 2.60 shows a typical beam type torque wrench which has an extension spanner
attached. If this combination is used to torque load a fastener then the following formula should
be used to calculate the wrench scale reading which corresponds to the specified torque value:

L
Scale reading = specified torque ×
L+X

Where L = distance between the driving tang and the centre of the handle
X = length of extension spanner between centres

Module 2.2 Mechanics 2-165


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Figure 2.60: A torque wrench fitted with an extension spanner

A simple way of calculating the scale reading required without using the formula is set out in the
following example, for which the specified torque loading is 300 lb in and the lengths of the
wrench and spanner are 10 and 5 inches respectively.

(a) Force required on wrench handle to produce a torque of 300 lb in is 300 lb in divided by
the distance between nut and wrench handle,

300 lb.in.
which is = 20 lb
10 in. + 5 in.

(b) Scale reading when force on handle is 20 lb is, 20 lb x 10 in 200 lb in.

Force must therefore be applied to the wrench handle until a reading of 200 lb in is shown on
the wrench scale, and this will represent a 300 lb in torque load applied to the nut. With the
‘break’ type wrench, the adjustment must be preset at 200 lb in.

NOTE: For the purpose of conversion, 1 lb.in. = 115 kg cm or 0.113 N.m.

When using an extension spanner with a torque wrench, the spanner and wrench
should be as nearly as possible in line. If it is necessary to diverge by more than 15°
from a straight line (due, for example, to intervening structure), then the direct
distance (D) between the nut and wrench handle must be substituted for ‘L + X’ in the
formula for calculating wrench scale reading. This is shown in figure 2.61, and the scale reading
in this instance will be equal to specified torque x

2-166 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Figure 2.61: A torque wrench fitted with an extension
spanner positioned out-of-line with the wrench

Whenever a torque wrench is used, it must be confirmed that the specified torque and the
wrench scale are in the same units; if not, then the specified torque should be converted, by
calculation, to the units shown on the wrench scale, and any measurements taken in
appropriate units.

When applying torque the wrench handle should be lightly gripped and force applied smoothly
at 90° to the axis of the wrench.

Couples
A ‘couple’ is a pair of forces of magnitude F that are equal and opposite but applied at points
separated by a distance d perpendicular to the forces. The combined moment of the forces
produces a torque Fd on the object on which they act.

An example is the cutting of an internal thread with a tap and tap wrench. The force applied at
one end of the wrench handle, multiplied by the distance to the centre of rotation is just half of
the torque felt at the tap itself, since there is an equal torque applied at the other wrench handle.

Torque applied by a couple

= one of the forces (F) x distance to centre of rotation (r) x 2

= one of the forces (F) x distance between the forces (d) = Fd

Another example is the forces applied to a car steering wheel.

Module 2.2 Mechanics 2-167


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-168 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Problems

1. Calculate the torque applied to a nut and bolt by a 12 in. spanner when a force of 12 lb.
is applied perpendicular at the end of the spanner.

2. How much force is required to torque a nut and bolt to 50 Nm with a wrench 0.5 m long?

3. A nut is to be torqued to 50 in.lb. A torque wrench of 17 in is used with an extension of 3


in. What setting should the torque wrench be adjusted to?

4. A ships wheel has a couple applied to it by the captain of 60 Nm. The diameter of the
wheel is 0.8 m. What is the force applied on just one side of the wheel?

Module 2.2 Mechanics 2-169


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-170 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Answers

1. 144 in.lb.
2. 100 N
3. 42.5 in.lb.
4. 75 N

Module 2.2 Mechanics 2-171


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-172 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
The Gyroscope
Gyros are fascinating to study and a great deal of material is available on them. For the most
part, we will be concerned with only two of the properties of spinning gyros. The first is the
tendency of a spinning gyro to remain fixed in space if it is not acted upon by outside forces
such as bearing friction. This is the property of rigidity. Rigidity is used to measure position in
position gyros such as the HSI (gyro compass) and ADI (artificial horizon).

The other property of a spinning gyro that concerns


concerns us is its right angle obstinacy. It never goes
in the direction that you push it, but off to one side. Figure 2.62 illustrates this obstinate
characteristic.

Figure 2.62: Force and resultant movement obstinacy – ‘Precession’

Whichever way you apply the force to the axis of a gyro, it will move in a direction 90o (in
direction of rotation) to the force. The speed at which it moves is proportional to the force
applied. This action is called precession. The force of procession is used in rate gyros, such
as those in a turn and slip indicator, where the speed of turn
turn is measured by the force that the
precessing gyro exerts on a spring.

Apparent Drift (or Wander)


Figure 2.63 illustrates the behavior of a gyro. A perfect gyro would be one without any external
forces acting upon it, mounted in a perfect suspension system
system that would give it complete
freedom of movement in all three axes. All the gyros in this figure are perfect gyros. Such gyros
are called free gyros.

Only four gyros are represented - A, B, C, and D. The other gyro symbols shown illustrate the
various positions of B, C, and D as the earth rotates.

Module 2.2 Mechanics 2-173


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Gyro A has its spin axis parallel with the spin axis of the earth, sitting on top of the North Pole. It
could maintain that position indefinitely.

Gyro B has its spin axis parallel to the earth’s spin axis,
axis, and is located above the equator. The
other gyros in its group represent Gyro B as it would appear at different times of the day. If we
were to look at Gyro B sitting on a table in front of us, we would see that the upper end of its
spin axis is pointing off toward the north star. As time goes on and the earth turns 360°, we
would not see any change in its attitude on the table. Its spin axis would always point toward the
north star.

Gyro C is situated on the equator. The other gyros in its group represent
represent Gyro C as it would
appear at different times of the day.

Let’s say that we have the Gyro C in front of us on a table. Its spin axis is parallel to the earth’s
surface. As time goes on and the earth rotates, we would see its spin axis gradually tilting
upward at one end until, six hours later (90° of ea rth rotation), we would see it perpendicular to
the earth’s surface, illustrated by the gyro shown to the right of the earth. Six hours later (behind
the earth out of sight in this drawing) the spin axis would once again be parallel to the earth, but
with the end which was first pointing east now pointing west.

Another six hours later,


the spin axis would once
again be perpendicular,
but this time the opposite
end of the axis would be
another six hours later.
When we get to the
same time of day at
which we started, the
gyro will again be
occupying its original
position.

Gyro D and its group


illustrate another chang-
ing aspect of a gyro, in
different positions as
viewed from the earth’s
surface at different times
of day.

These perfect gyros


illustrate what any gyro
tries to do but cannot
because of its orientation
of the spin axis - always in Figure 2.63: Apparent Drift
the same direction in space.

2-174 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Transport Drift (or Wander)
The outer ring of gyros in Figure 2.64 demonstrates that a completely free gyro in an aircraft
circling the earth would be perpendicular to the earth’s surface at only two points.

The gyros drawn in the aircrafts are continuously being corrected to a vertical position as the
aircraft moves around the surface of the earth. The corrections are gentle and slow, since the
amount of correction needed in a ten minute period, for example, is small The gyro is relatively
very stable during the pitch and roll maneuvers of the aircraft. Such a gyro is called an earth
gyro or tied gyro.

The gyro’s stable position with respect to the movements of the aircraft makes it possible for the
pilot to know the actual attitude of his aircraft, nose up or down, and wings level or not. This is
quite important to him when all he can see out of the window is a gray fog.

The aircraft attitude information derived from the gyros is also used by such systems as the
autopilot, radar antenna stabilization, flight recorders and flight directors.

Figure 2.64: Transport Drift

Module 2.2 Mechanics 2-175


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Figure 2.65: Horizon indication and Compass indication (Position Gyros)

Figure 2.66: Turn Rate indication (Rate Gyro)

2-176 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Fluid Dynamics
The Atmosphere
On November 21, 1783, a hot air balloon carrying Marquis d’Arlandes, flew 5 miles across the
skies of Paris opening up new possibilities in travel and a fresh interest in our atmosphere. It
wasn’t however, until heavier-than-air flight became a reality that a detailed understanding of
the medium enveloping our globe became essential.

The atmosphere is a mixture of gases which we call air. Dry air is composed of approximately
21% oxygen, 78% nitrogen, and 1% carbon dioxide. These percentages remain fairly constant
as we ascend in altitude. However, the density of air decreases. This drop in density with
altitude has great significance in aviation as it not only places limits on the attainable altitudes,
but also the powerplant performance of an aircraft.

The mapping out of our atmosphere, that is, determining its density, pressure, and temperature
at different altitudes, required the effort of many individuals working over many years. The fruit
of this labour is a vast quantity of data which has led to the definition of a standard atmosphere.
The standard atmosphere, a term coined by Willis Ray Gregg in 1922, is a compilation of mean
annual atmospheric properties. Since our atmosphere undergoes seasonal variations in
properties such as temperature, a mean or average value is used. Tables 2.6 and 2.7 are two
tables of values for the standard atmosphere. The first table (table 2.6) gives values in English
units and the second (table 2.7) in metric units. It must be kept in mind that the numbers in
these tables are annual averages which can be useful for reference purposes but do not
indicate the actual atmospheric conditions existing at any particular moment.

EXAMPLE:

Using the Gas Law and the temperature and pressure at an altitude of 12,000 meters listed in
the Standard Atmosphere Table (table 2.5), verify that the density of air at this altitude is 0.312
kg/m3.
We will use the equation
P
ρ=
RT

19.4 kPa
ρ=
(287 Pa m 3 /kg K)(217 K)

= 0.312 kg/m3

Module 2.2 Mechanics 2-177


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Cabin Altitude
Cabin altitude is a term used to express cabin pressure in terms of equivalent altitude above
sea level. For example, a cabin altitude of 6,000 feet means that the pressure inside the aircraft
cabin is the same as the atmospheric pressure at an altitude of 6,000 feet. Looking at the
Standard Atmosphere Table (table 2.6), the pressure is found to be 1,696 lbs/ft2 which upon
division by 144 gives the pressure in lb/in2 to be 11.78 lbs/in2

At a cabin altitude of 8,000 feet, the passengers and crew can ride in relative comfort without
any special oxygen supply. Planes which fly at much higher altitudes than 8,000 feet must be
furnished with a special atmosphere control system. It is highly advantageous to fly at high
altitudes both for economy of fuel consumption, and the smooth air high above the level of
turbulent weather systems. At these high altitudes, the pressure outside the plane can be
significantly lower than the cabin pressure.

At 8,000 ft., the Standard Atmosphere Table tells us that the air pressure is 1,572 lbs/ft2 or
10.92 lbs/in2 This is the pressure that is normally maintained in the cabin even though the plane
is flying at a higher altitude.

Suppose the plane is flying at an altitude of 40,000 ft. At this altitude the pressure (from the
table 5-1) is 393 lbs/ft2 or 2.73 lbs/in2 This means that for a cabin altitude of 8,000 ft. for a plane
flying at 40,000 ft., there is a net outward pressure of 8.19 lbs/in2 This number was obtained by
subtracting 2.73 lbs/in2 from 10.92. For a Learjet with a pressurized area of 45,000 in2, we are
dealing with a bursting force of over 368,000 lbs. (8.19 x 45 thousand). in addition to being able
to withstand this much force, a safety factor of 1.33 is generally used by design engineers.
Therefore, the pressurized portion of the fuselage must be constructed to have an ultimate
strength of over 460 thousand pounds or about 230 tons. The challenge of finding lightweight
materials which can withstand these large forces is great.

In the description of an aircraft’s air conditioning and pressurization system, a differential


pressure is given. The differential pressure is the maximum difference between cabin pressure
and atmospheric pressure which the pressurization system can sustain. For example, the air-
cycle air-conditioning system of a Boeing 747 can maintain a pressure differential of 8.9 lbs/in2
This means that the system can maintain a cabin pressure 8.9 lbs/in2 greater than the
atmospheric pressure surrounding the plane. This also means that there is an upper limit
imposed by the pressurization system on the altitude at which the plane can fly.

2-178 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Standard Atmosphere — English Units

Altitude Temperature Pressure Density


(ft.) (°R) (lb./Ft2) (Slug/Ft3)

0 519 2,116 0.002377


500 517 2,078 0.002342
1,000 515 2,041 0.002308
2,000 512 1,968 0.002241
3,000 508 1,897 0.002175
4,000 504 1,828 0.002111
5,000 501 1,761 0.002048
6,000 497 1,696 0.001987
7,000 494 1,633 0.001927
8,000 490 1,572 0.001869
9,000 487 1,513 0.001811
10,000 484 1,456 0.001756
15,000 465 1,195 0.001496
20,000 447 973 0.001267
25,000 430 786 0.001066
30,000 412 630 0.000891
35,000 394 499 0.000738
40,000 390 393 0.000585
45,000 390 309 0.000462
50,000 390 244 0.000364
55,000 390 192 0.000287
60,000 390 151 0.000226
65,000 390 119 0.000178

Table 2.6: Quantities within the Standard Atmosphere (English Units)

Module 2.2 Mechanics 2-179


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Standard Atmosphere — Metric Units

Altitude Temperature Pressure Density


(M) (K) (kPa) (kg/m3)

0 288 101.3 1.225


100 288 100.1 1.213
200 287 98.9 1.202
300 286 97.8 1.190
400 286 96.6 1.179
500 285 95.5 1.167
600 284 94.3 1.156
700 284 93.2 1.145
800 283 92.1 1.134
900 282 91.0 1.123
1,000 282 89.9 1.112
1,500 278 84.6 1.058
2,000 275 79.5 1.007
2,500 272 74.7 0.957
3,000 269 70.1 0.909
3,500 265 65.8 0.863
4,000 262 61.7 0.819
4,500 259 57.8 0.777
5,000 256 54.0 0.736
5,500 252 50.5 0.697
6,000 249 47.2 0.660
6,500 246 44.1 0.624
7,000 243 41.1 0.590
7,500 239 38.3 0.557
8,000 236 35.7 0.526
8,500 233 33.2 0.496
9,000 230 30.8 0.467
9,500 227 28.6 0.439
10,000 223 26.5 0.414
12,000 217 19.4 0.312
14,000 217 14.2 0.228
16,000 217 10.4 0.166
18,000 217 7.57 0.122
20,000 217 5.53 0.0889
22,000 217 4.04 0.0650
Table 2.7: Quantities within the Standard Atmosphere (Metric Units)

2-180 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Humidity
Some water in the form of invisible vapour is intermixed with the air throughout the atmosphere.
It is the condensation of this vapour which gives rise to most weather phenomena: clouds, rain,
snow, dew, frost and fog. There is a limit to how much water vapour the air can hold and this
limit varies with temperature. When the air contains the maximum amount of vapour possible for
a particular temperature, the air is said to be saturated. Warm air can hold more vapour than
cold air. In general the air is not saturated, containing only a fraction of the possible water
vapour.

The amount of vapour in the air can be measured in a number of ways. The humidity of a
packet of air is usually denoted by the mass of vapour contained within it, or the pressure that
the water vapour exerts. This is the absolute humidity of air. Relative humidity is measured
by comparing the actual mass of vapour in the air to the mass of vapour in saturated air at the
same temperature. For example, air at 10°C contains 9.4 g/m3 (grams per cubic metre) of water
vapour when saturated. If air at this temperature contains only 4.7 g/m3 of water vapour, then
the relative humidity is 50%.

When unsaturated air is cooled, relative humidity increases. Eventually it reaches a


temperature at which it is saturated. Relative humidity is 100%. Further cooling leads to
condensation of the excess water vapour. The temperature at which condensation sets in is
called the dew point. The dew point, and other measures of humidity can be calculated from
readings taken by a hygrometer. A hygrometer has two thermometers, one dry bulb or standard
air temperature thermometer, and one wet bulb thermometer. The wet bulb thermometer is an
ordinary thermometer which has the bulb covered with a muslin bag, kept moist via an
absorbent wick dipped into water. Evaporation of water from the muslin lowers the temperature
of the thermometer. The difference between wet and dry bulb temperatures is used to calculate
the various measures of humidity.

Definitions

Absolute humidity: The mass of water vapour in a given volume of air (i.e., density of water
vapour in a given parcel), usually expressed in grams per cubic meter

Actual vapour pressure: The partial pressure exerted by the water vapour present in a parcel.
Water in a gaseous state (i.e. water vapour) exerts a pressure just like the atmospheric air.
Vapour pressure is also measured in Millibars.

Condensation: The phase change of a gas to a liquid. In the atmosphere, the change of water
vapour to liquid water.

Dewpoint: the temperature air would have to be cooled to in order for saturation to occur. The
dewpoint temperature assumes there is no change in air pressure or moisture content of the air.

Module 2.2 Mechanics 2-181


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Dry bulb temperature: The actual air temperature. See wet bulb temperature below.

Freezing: The phase change of liquid water into ice.

Evaporation: The phase change of liquid water into water vapour.

Melting: The phase change of ice into liquid water.

Mixing ratio: The mass of water vapour in a parcel divided by the mass of the dry air in the
parcel (not including water vapour).

Relative humidity: The amount of water vapour actually in the air divided by the amount of
water vapour the air can hold. Relative humidity is expressed as a percentage and can be
computed in a variety of ways. One way is to divide the actual vapour pressure by the saturation
vapour pressure and then multiply by 100 to convert to a percent.

Saturation of air: The condition under which the amount of water vapour in the air is the
maximum possible at the existing temperature and pressure. Condensation or sublimation will
begin if the temperature falls or water vapour is added to the air.

Saturation vapour pressure: The maximum partial pressure that water vapour molecules
would exert if the air were saturated with vapour at a given temperature. Saturation vapour
pressure is directly proportional to the temperature.

Specific humidity: The mass of water vapour in a parcel divided by the total mass of the air in
the parcel (including water vapour).

Sublimation: In meteorology, the phase change of water vapour in the air directly into ice or the
change of ice directly into water vapour. Chemists, and sometimes meteorologists, refer to the
vapour to solid phase change as "deposition."

Wet bulb temperature: The lowest temperature that can be obtained by evaporating water into
the air at constant pressure. The name comes from the technique of putting a wet cloth over the
bulb of a mercury thermometer and then blowing air over the cloth until the water evaporates.
Since evaporation takes up heat, the thermometer will cool to a lower temperature than a
thermometer with a dry bulb at the same time and place. Wet bulb temperatures can be used
along with the dry bulb temperature to calculate dew point or relative humidity.

2-182 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Problems

1. Verify, that using the Gas Law (ρ = P/RT) and the temperature and pressure from the
Standard Atmosphere Table, at an altitude of 65,000 ft., the density of air is 0.000178
slug/ft.3

2. A pressurized Cessna Centurion II has a cabin pressurization system which can maintain
a pressure differential of 3.45 lbs/in2 What is the maximum altitude at which the plane
can fly and still maintain a cabin altitude of 8,000 feet?

(Hint: convert the of 3.45 lbs/in2 to lbs/ft2 and compare with the Standard Atmosphere
table)

3. What is the maximum altitude at which this same Cessna plane can fly and maintain a
cabin altitude of 6,000 ft?

Module 2.2 Mechanics 2-183


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-184 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Answers

2. About 18,000 ft
3. About 14,000 ft

Module 2.2 Mechanics 2-185


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-186 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Density and Specific Gravity

Density
The density of a material is defined as the mass of a sample of the material divided by the
volume of the same sample. The symbol used for density is the Greek letter rho, (ρ).

m
ρ=
V
Other algebraic forms of this same equation are:

m
m = ρV and V=
ρ
Density is a very important and useful concept. If a body is made of a certain kind of material its
density is known. If the weight of the body is also known, it is possible to determine the volume
of this body. Similarly, if the kind of material and volume are known it is possible to determine
the weight of the body.

Table 2.8 is a table of densities. You can refer to this table when you solve the problems
dealing with mass, weight, and volume.

Module 2.2 Mechanics 2-187


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Densities of Liquids and Solids

Liquids Kg/m3 Slug/ft3 Metals Kg/m3 Slug/ft3


Water 1000 1.940 Aluminium 2700 5.25
Sea Water 1030 2.00 Cast Iron 7200 14.0
Benzine 879 1.71 Copper 8890 17.3
Alcohol 789 1.53 Gold 19300 37.5
Gasoline 680 1.32 Lead 11340 22.0
Kerosene 800 1.55 Nickel 8850 17.2
Sulphuric Acid 1831 3.55 Silver 10500 20.4
Mercury 13600 26.3 Steel 7800 15.1
Tungsten 19000 37.0
Non-Metals Zinc 7140 13.9
o o
Ice (32 F, 0 C) 922 1.79 Brass 8700 16.9
Concrete 2300 4.48

Glass 2,600 4.97 Woods


Granite 2700 5.25 Balsa 130 0.25
Pine 480 0.93
Maple 640 1.24
Oak 720 1.4
Ebony 1200 2.33
Table 2.8: Comparison of densities – Liquids and solids

2-188 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
EXAMPLE:
An order has been placed for 120 gal. of lubricating oil. How much will this oil weigh?

1ft 3
V = 120 gal × = 16.0 ft 3
7.481 gal

slug
ρ = 1.75
ft 3

The density of the lubricating oil has been obtained from table 2-1.

m = ρV

W 150 lbs
m= = = 4.69 slug
g 32 lbs/slug

 32 lbs 
W = mg = (28 slugs)   = 896 lbs
 slug 

EXAMPLE:

An order has been placed for 150 lbs. of turpentine. How many gallons of turpentine will be
delivered?

W 150 lbs
m= = = 4.69 slug
g 32 lbs/slug

m 4.69 slug
V= = 3
= 2.78 ft 3
ρ 1.69 slug/ft

7.481 gal
V = 2.78 ft 3 × = 20.8 gal
1ft

Module 2.2 Mechanics 2-189


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Specific Gravity
The term “specific gravity” is closely related to the idea of density. The definition is as follows:

density of the substance


Specific Gravity =
density of water

The calculation will give the same result (for a given substance) no matter what units are used.
The example below will calculate the specific gravity of sulphuric acid (see table 2.8).

If we use the metric units (kg/m3) we obtain:

1,831
Specific Gravity = = 1 .83
1,000

If we use the English units (slug/ft3) we obtain:

3.55
Specific Gravity = = 1.83
1.94

The specific gravity number (1.83) is unitless. It tells us that, for sulphuric acid, the density is
1.83 times as dense as water.

2-190 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Problems

1. What is the specific gravity of kerosene?

2. What is the specific gravity of aluminium?

3. What is the specific gravity of ice?

4. What is the specific gravity of glass?

5. What is the weight of 85 gallons of kerosene?

Module 2.2 Mechanics 2-191


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-192 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Answers

1. 0.8
2. 2.7
3. 0.922
4. 2.6
5. 544 lbs. or 3029 N (Hint: Calculate weight of water, 1 litre = 1kg, or 1 pint = 1 lb, then
convert to kerosene by multiplying by its specific gravity of 0.8)

Module 2.2 Mechanics 2-193


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-194 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Compressibility in Fluids
Fluids are defined as any substance which flows readily. Gases and liquids are such
substances

Gases
A gas is relatively easy to compress, and the effects of which have already been discussed in
the chapter on Pressure and Force.

Liquids
Many people think that a liquid is incompressible. However, liquids are, like any material, to a
certain amount compressible. In calculations, the amount of compressibility of liquid is
considered to be 1 volume-% per 100 bar. This means that for example when there is liquid
supplied to a 200 litre oil drum which already is completely filled with liquid, the pressure
increases with 100 bar for each 2 litre of extra supplied liquid. When we supply 3 litre of extra oil
the pressure increases with 150 bar. The compressibility of liquid plays a key role in for example
fast hydraulic systems like servo-systems of a flight simulator. To obtain a maximum dynamic
performance, the compressibility should be as little as possible. This is achieved by mounting
the control valves directly on the hydraulic motor or cylinder. In that case the amount of liquid
between the control valve and the motor/cylinder is minimised.

In some situations, the compressibility of liquids is made use of in design. A ‘liquid spring’ for
example, is the principle of a particular type of landing gear leg, which uses no gas. The leg is
completely filled with oil. The oil is compressed under the extremely large forces encountered
on landing, and the shock of the landing is absorbed by the compressibility of the liquid.

Module 2.2 Mechanics 2-195


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-196 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Viscosity
Viscosity is a measure of the resistance of a fluid to being deformed by either shear stress or
extensional stress. It is commonly perceived as "thickness", or resistance to flow. Viscosity
describes a fluid's internal resistance to flow and may be thought of as a measure of fluid
friction. Thus, water is "thin", having a lower viscosity, while vegetable oil is "thick" having a
higher viscosity. All real fluids (except superfluids) have some resistance to stress, but a fluid
which has no resistance to shear stress is known as an ideal fluid or inviscid fluid.

The study of viscosity is known as rheology.

Viscosity coefficients
When looking at a value for viscosity, the number that one most often sees is the coefficient of
viscosity. There are several different viscosity coefficients depending on the nature of applied
stress and nature of the fluid.

• Dynamic viscosity determines the dynamics of an incompressible fluid;


• Kinematic viscosity is the dynamic viscosity divided by the density;
• Volume viscosity determines the dynamics of a compressible fluid;
• Bulk viscosity is the same as volume viscosity

Shear viscosity and dynamic viscosity are much better known than the others. That is why they
are often referred to as simply viscosity. Simply put, this quantity is the ratio between the
pressure exerted on the surface of a fluid, in the lateral or horizontal direction, to the change in
velocity of the fluid as you move down in the fluid (this is what is referred to as a velocity
gradient). For example, at "room temperature", water has a nominal viscosity of 1.0 × 10-3 Pa·s
and motor oil has a nominal apparent viscosity of 250 × 10-3 Pa·s.

Viscosity Measurement
Dynamic viscosity is measured with various types of viscometer. Close temperature control of
the fluid is essential to accurate measurements, particularly in materials like lubricants, whose
viscosity can double with a change of only 5°C. For some fluids, it is a constant over a wide
range of shear rates. These are Newtonian fluids.

The fluids without a constant viscosity are called Non-Newtonian fluids. Their viscosity cannot
be described by a single number. Non-Newtonian fluids exhibit a variety of different correlations
between shear stress and shear rate.

One of the most common instruments for measuring kinematic viscosity is the glass capillary
viscometer.

In paint industries, viscosity is commonly measured with a Zahn cup, in which the efflux time is
determined and given to customers. The efflux time can also be converted to kinematic
viscosities (cSt) through the conversion equations.

Also used in paint, a Stormer viscometer uses load-based rotation in order to determine
viscosity. The viscosity is reported in Krebs units (KU), which are unique to Stormer
viscometers.

Module 2.2 Mechanics 2-197


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Vibrating viscometers can also be used to measure viscosity. These models use vibration rather
than rotation to measure viscosity.

Units of Measure

Viscosity (dynamic/absolute viscosity)


Dynamic viscosity and absolute viscosity are synonymous. The symbol for viscosity is the Greek
symbol eta (η), and dynamic viscosity is also commonly referred to using the Greek symbol mu
(µ). The SI physical unit of dynamic viscosity is the pascal-second (Pa·s), which is identical to
1 kg·/(m s) (kilogram per metre-second). If a fluid with a viscosity of one Pa·s is placed between
two plates, and one plate is pushed sideways with a shear stress of one pascal, it moves a
distance equal to the thickness of the layer between the plates in one second.

Kinematic viscosity
In many situations, we are concerned with the ratio of the viscous force to the inertial force, the
latter characterised by the fluid density ρ. This ratio is characterised by the kinematic viscosity
(ν), defined as follows:

where µ is the (dynamic) viscosity, and ρ is the density.

Kinematic viscosity (Greek symbol: ν) has SI units (m2·/ s). The cgs physical unit for kinematic
viscosity is the stokes (abbreviated S or St), named after George Gabriel Stokes. It is
sometimes expressed in terms of centistokes (cS or cSt). In U.S. usage, stoke is sometimes
used as the singular form.

1 stokes = 100 centistokes = 1 cm2/s = 0.0001 m2/s.


1 centistokes = 1 mm2/s

Viscosity of air
The viscosity of air depends mostly on the temperature. At 15.0 °C, the viscosity of air is
1.78 × 10−5 kg/(m·s).

Viscosity of water
The viscosity of water is 8.90 × 10−4 Pa·s at about 25 °C.

2-198 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Drag and Streamlining
In fluid dynamics, drag (sometimes called resistance) is the force that resists the movement of a
solid object through a fluid (a liquid or gas). Drag is made up of friction forces, which act in a
direction parallel to the object's surface (primarily along its sides, as friction forces at the front
and back cancel themselves out), plus pressure forces, which act in a direction perpendicular to
the object's surface. For a solid object moving through a fluid or gas, the drag is the sum of all
the aerodynamic or hydrodynamic forces in the direction of the external fluid flow. (Forces
perpendicular to this direction are considered lift). It therefore acts to oppose the motion of the
object, and in a powered vehicle it is overcome by thrust.

Types of drag are generally divided into three categories: profile drag (called ‘parasitic’ drag in
USA), lift-induced drag (also known as vortex drag or induced drag), and wave drag. Profile
drag includes form drag, skin friction, and interference drag. Lift-induced drag is only relevant
when wings or a lifting body are present, and is therefore usually discussed only in the aviation
perspective of drag. Wave drag occurs when a solid object is moving through a fluid at or near
the speed of sound in that fluid. The overall drag of an object is characterized by a
dimensionless number called the drag coefficient, and is calculated using the drag equation.
Assuming a constant drag coefficient, drag will vary as the square of velocity. Thus, the
resultant power needed to overcome this drag will vary as the cube of velocity. The standard
equation for drag is one half the coefficient of drag multiplied by the fluid density, the cross
sectional area of the specified item, and the square of the velocity.

Wind resistance or air resistance is a layman's term used to describe drag. Its use is often
vague, and is usually used in a relative sense (e.g., A badminton shuttlecock has more wind
resistance than a squash ball).

Stokes's Drag
The equation for viscous resistance or linear drag is appropriate for small objects or particles
moving through a fluid at relatively slow speeds where there is no turbulence. In this case, the
force of drag is approximately proportional to velocity, but opposite in direction. The equation for
viscous resistance is:

Viscous resistance = - bv

where:
b is a constant that depends on the properties of the fluid and the dimensions of
the object, and
v is the velocity of the object.

For the special case of small spherical objects moving slowly through a viscous fluid, George
Gabriel Stokes derived an expression for the drag constant,

B = 6π η r
where:
r is the Stokes radius of the particle, and η is the fluid viscosity.

Module 2.2 Mechanics 2-199


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
For example, consider a small sphere with radius r = 0.5 micrometre (diameter = 1.0 µm)
moving through water at a velocity v of 10 µm/s. Using 10−3 Pa·s as the dynamic viscosity of
water in SI units, we find a drag force of 0.28 pN. This is about the drag force that a bacterium
experiences as it swims through water.

Drag Coefficient
The drag coefficient (Cd) is a dimensionless quantity that describes a characteristic amount of
aerodynamic drag caused by fluid flow, used in the drag equation. Two objects of the same
frontal area moving at the same speed through a fluid will experience a drag force proportional
to their Cd numbers. Coefficients for rough unstreamlined objects can be 1 or more, for smooth
objects much less.

Aerodynamic drag = Cd ½ ρ V2 A

Where
Cd = drag coefficient (dimensionless)
ρ = fluid density (slug/ft3 or kg/m3)
V = Velocity of object (ft/s or m/s)
A = projected frontal area (ft2 or m2)

The drag equation is essentially a statement that, under certain conditions, the drag force on
any object is approximately proportional to the square of its velocity through the fluid.

Figure 2.67: Effect of airflow on a flat plate

2-200 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
A Cd equal to 1 would be obtained in a case where all of the fluid approaching the object is
brought to rest, building up stagnation pressure over the whole front surface. Figure 2.67 (top)
shows a flat plate with the fluid coming from the right and stopping at the plate. The graph to the
left of it shows equal pressure across the surface. In a real flat plate the fluid must turn around
the sides, and full stagnation pressure is found only at the centre, dropping off toward the edges
as in the lower figure and graph. The Cd of a real flat plate would be less than 1, except that
there will be a negative pressure (relative to ambient) on the back surface.

Some examples of Cd
Object Cd
A smooth brick 2.1
A bicycle plus rider 0.9
A rough sphere 0.4
A smooth sphere 0.1
A flat plate parallel to the flow 0.001
A bullet 0.295
A man (upright position) 1.0-1.3
A flat plate perpendicular to flow 1.28
A skier 1.0-1.1
Wires and cables 1.0-1.3

Drag coefficients of some complete aircraft


Aircraft type Cd
Cessna 172/182 0.027
Cessna 310 0.027
Learjet 24 0.022
Boeing 747 0.031
X-15 0.95

Streamlining
Streamlining is the shaping of an object, such as an aircraft body or wing, to reduce the amount
of drag or resistance to motion through a stream of air. A curved shape allows air to flow
smoothly around it. A flat shape fights air flow and causes more drag or resistance. Streamlining
reduces the amount of resistance and increases lift.

To produce less resistance, the front of the object should be well rounded and the body should
gradually curve back from the midsection to a tapered rear section.

Figure 2.68 shows how the drag of a flat plate can be reduced if its shape is changed to a
sphere, and more still if it is streamlined with fairings.

Module 2.2 Mechanics 2-201


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Figure 2.68: Streamlining of an object reduces its drag

2-202 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Bernoulli’s Principle

Basic Definitions
Before we begin our discussion of the lift and drag on an aircraft wing, the following definitions
must be understood.

The Pitot tube (named after Henri Pitot in 1732) measures a fluid velocity by converting the
kinetic energy of the flow into potential energy. The conversion takes place at the stagnation
point, located at the Pitot tube entrance (Figure 2.69). A pressure higher than the free-stream
(i.e. dynamic pressure) results from this conversion.

This static pressure is measured at the static taps (known as static ports or vents). The static
pressure is not affected by the speed of the aircraft, but is dependant upon the surrounding
atmospheric static pressure.

Pitot Pressure is the sum of static and dynamic pressures, thus:

Pitot Pressure = Static Pressure + Dynamic Pressure

Figure 2.69: A combined Pitot tube and static taps

Module 2.2 Mechanics 2-203


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Bernoulli’s Principle applies the ideas of work and energy and the conservation of energy to a
mass of fluid (liquid or gas). Since it is not as easy to think of a mass of fluid as it is to think of a
discrete body, the derivation of this principle requires some thought and effort.

It is worth the thought and effort, however, since this principle is the basic principle of the flight
of heavier-than-air aircraft.

We review that the density of a fluid (ρ) is related to the mass and volume of the sample of fluid
by the relation:

m = ρV

Figure 2.70: A volume of fluid

Figure 2.71: Fluid flowing through a tube of increasing cross sectional area

The fluid flows from a region where the cross-sectional area is less (labelled with 1's in the di-
agram) to a region where the cross-sectional area is greater (labelled with 2’s in the diagram).
We assume that the volume of fluid in the left cylindrical shape of fluid (labelled with 1’s) is
equal to the volume of fluid in the right cylindrical shape of fluid (labelled with 2’s).

Hence, the volume flow rate in any part of the tube is constant, regardless of the tube diameter.
And, since the density of the fluid is constant (and air flowing at subsonic speed is considered
incompressible) than the mass flow rate is also constant, regardless of the diameter of the tube.

2-204 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
The Venturi Tube
A venturi tube is a tube constructed in such a way that the cross sectional area of the tube
changes from a larger area to a smaller area and finally back to the same larger area. As a fluid
flows through this tube the velocity of the fluid changes from a lower velocity to a higher velocity
and finally back to the same lower velocity. We note that, if the rate (volume per second) of fluid
flow is to remain constant, the fluid must flow faster when it is flowing through the smaller area.

A diagram of a venturi tube is shown in figure 2.72.

1 2

Figure 2.72: A venturi tube

The height of the fluid column in the vertical tubes at the three places shown in the diagram, is
an indication of the fluid pressure. As we expect from Bernoulli’s Principle, the pressure is
greater where the velocity is lower and vice versa.

Venturi tubes in different shapes and sizes are often used in aircraft systems.

If we consider the types of energy involved in the flowing fluid, we find that there are three types
– potential (gravitational), pressure and kinetic energies.

Now if we consider only two positions in the venturi – the wide part (marked ‘1’) and the narrow
part (the throat, marked ‘2’), and consider the conservation of energy principle, we have:

Potential Energy at 1 Potential Energy at 2


+ +
Pressure Energy at 1 = Pressure Energy at 2
+ +
Kinetic energy at 1 Kinetic energy at 2

The above is assumed since the total energy in the fluid cannot change, only transferred from
one form to another. This then, is the basis for Bernoulli’s Formula.

Module 2.2 Mechanics 2-205


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Since the venturi in this case is horizontal, there in no change in potential energy, and so the
potential energies can be cancelled from the formula thus:

Pressure Energy at 1 Pressure Energy at 2


+ = +
Kinetic energy at 1 Kinetic energy at 2

1
Since Kinetic Energy is mV 2 where m = mass of fluid, and V = velocity of fluid
2
P
and Pressure energy is m where P = pressure, ρ = density of the fluid
ρ

Thus:

P1 1 2 P2 1 2
m + mV1 = m + mV2
ρ 2 ρ 2

Note that the mass, m, has no suffix, since the mass flow rate is constant regardless of the area
of the flow. The density, ρ, is also a constant since the fluid is considered incompressible (even
air, providing its velocity is subsonic).

Cancelling the mass, m, from each equation, and multiplying each term by the density, ρ, gives
us

P1 + ½ ρ V12 = P2 + ½ ρ V22

This is the standard mathematical form of the Bernoulli’s Equation. It can be rearranged to give
the pressure difference (for example between the upper and lower surfaces of a wing) thus:

P1 – P2 = ½ ρ V22 - ½ ρ V12

Factorising gives:
P 1 – P2 = ½ ρ (V22 - V12)

2-206 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Application of Bernoulli’s Principle to Aerofoil Sections

The relative wind direction is the direction of the


airflow with respect to the wing and is opposite to
the path of flight (figure 2.73).

The chord line of a wing is a straight line con-


necting the leading edge of a wing to its trailing
edge (figure 2.74).
Figure 2.73: Relative wind

Figure 2.74: Chord line


The angle of attack
Angle of Attack is the angle between the chord line
of a wing and the relative wind direction (figure 2.75).

Figure 2.76 shows the cross section of a wing at rest


and subject to atmospheric pressure which on the
average is 14.7 lbs./in.2

Figure 2.75:
Angle of attack

A force of 14.7 lbs. can be imagined as acting perpendicular


to every square inch of the wing. The resultant of these 14.7
lbs. force vectors is zero and therefore has no effect on the
dynamics of the plane.

Figure 2.76: Pressure forces on


an aerofoil

It is the motion of air past the wing that alters the pressure pattern. Whether the wing is in
motion through the air or the air is flowing past a stationary wing the result is the same.

For example, if a plane is moving through stationary air at a speed of 200 MPH, the effect is the
same (as far as the plane and air are concerned) as if the plane were stationary and the air was
moving with velocity 200 MPH past the plane.

There is a thin layer of air in direct contact with the wing surface, which, due to skin friction, is
actually stationary (relative to the wing). This is called the boundary layer. In these
discussions we will disregard the boundary layer and assume that the airflow is unaffected by
friction.

Module 2.2 Mechanics 2-207


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
As air streams past the wing of a plane, the speed of the air past the upper surface of the wing
is greater than the speed of the air past the lower surface of the wing.

These exact speeds are determined by the shape of the wing and the angle of attack.

For example, if the speed of the relative wind (equal to the speed of the plane) is 200 MPH, the
speed of the air past the upper surface of the wing may be 210 MPH and the speed of air past
the lower surface of the wing may be 195 MPH. As indicated above, the exact values for a given
case depend on the shape of the wing and the angle of attack.

In this example, we could say that the speed past the upper surface of the wing is [1.05 (200
MPH)] and the speed past the lower surface if the wing is [0.975 (200 MPH)].

In figure 2.77, the following symbols apply: V1

P1 = pressure on the upper surface of the wing


P2 = pressure on the lower surface of the wing
Vo = relative wind velocity V2
V1 = Wind velocity over upper surface
V2 = Wind velocity over lower surface Figure 2.77: Velocities and pressures
ρ = density of the air above and below an aerofoil

We apply Bernoulli’s principle

P1 + ½ ρ V12 = P2 + ½ ρ V22

We note that the ones refer to the upper surface and the twos apply to the lower surface of the
wing.

P1 + ½ ρ V12 = P2 + ½ ρ V22

½ ρ V12 - ½ ρ V22 = P2 - P1

½ ρ (V12 - V22) = P2 - P1

When finding the lift on a wing, the pressure difference between the upper and lower surfaces is
found from the above equation, and, since Force = Pressure x Area, simply multiply the
calculated pressure difference by the area of the wing, thus:

∆P = ½ ρ (V22 - V12), and Lift (or Force) = ∆P x Area

Lift = ∆P x Area

(Note: In some questions, the weight of the aircraft will be quoted. Thus, if the aircraft is flying
straight and level, the Lift = Weight).

2-208 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Problems

1. An aeroplane having wing area 500 ft2 is moving at 300 ft/sec. The speed of the air
moving past the top surface of the wing is 400 ft/sec. and the speed of the air past the
bottom surface of the wing is 200 ft/sec. The density of the air is 0.0025 slug/ft3. What is
the lift?

2. An aeroplane having wing area 400 ft2 is cruising at 230 ft/sec. The speed of the air
moving past the top surface of the wing is 240 ft/sec and the speed of the air past the
bottom surface of the wing is 230 ft/sec. The density of the air is 0.0025 slug/ft3. What is
the weight of the aeroplane?

3. An aeroplane is cruising at 310 ft/sec. The speed of the air moving past the top surface
of the wing is 340 ft/sec and the speed of the air past the bottom surface of the wing is
300 ft/sec. The density of the air is 0.001 slug/ft3 The weight of the aeroplane is 12,800
lbs. What is the wing area?

Module 2.2 Mechanics 2-209


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-210 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012
Answers

1. 75,000 lbs.
2. 2350 lbs.
3. 1000 ft2

Module 2.2 Mechanics 2-211


Use and/or disclosure is
governed by the statement For Training Purposes Only
on page 2 of this chapter. ST Aerospace Ltd
© Copyright 2012
Intentionally Blank

2-212 Module 2.2 Mechanics


Use and/or disclosure is
For Training Purposes Only governed by the statement
ST Aerospace Ltd on page 2 of this chapter.
© Copyright 2012

You might also like