Sei sulla pagina 1di 299

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

net/publication/283293004

Theoretical Mechanics. Theory and Applications

Book · November 2012


DOI: 10.13140/RG.2.1.1786.1842

CITATIONS READS

2 5,389

1 author:

Dumitru Deleanu
Maritime University of Constanta
62 PUBLICATIONS   56 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

New indicators of chaos and regularity View project

All content following this page was uploaded by Dumitru Deleanu on 29 October 2015.

The user has requested enhancement of the downloaded file.


DUMITRU DELEANU

THEORETICAL MECHANICS

Theory and applications

Editura
NAUTICA
I
II
DUMITRU DELEANU

THEORETICAL MECHANICS

Theory and applications

Editura
NAUTICA

III
Copyright © 2012, Editura NAUTICA
Toate drepturile asupra acestei ediŃii sunt rezervate Editurii

Editura NAUTICA
Editură recunoscută de CNCSIS
Str. Mircea cel Bătrân nr. 104
900663 ConstanŃa, România
tel.: +40-241-66.47.40
fax: +40-241-61.72.60
e-mail: info@imc.ro
www.edituranautica.org.com

Descrierea CIP a Bibliotecii NaŃionale a României:

DELEANU, DUMITRU
Theoretical mechanics. Theory and applications / Dumitru Deleanu –
ConstanŃa: Nautica, 2012

Bibliogr.
ISBN 978-606-8105-89-5

531

IV
Foreword

Mechanics is one of the fundamental nature sciences with a very wide research
scope and it is concerned with the mechanical motion of the material systems. In time, the
mechanics differentiated into three distinct branches: classical mechanics, relativistic
mechanics, and the quantum mechanics.
The classical mechanics (Newtonian), having the Italian scientist Galileo Galilees
(1564-1642) and the British physicist, mathematician and astrologer Isaac Newton (1642-
1727) as founders, comprises the laws of the motion of the macro objects the speed of
which is small as compared to the speed of the electromagnetic waves in vacuum. The
Newtonian mechanics is considered as the first theoretical science of the nature. It is
divided in its turn into two branches: the classical mechanics and the applied mechanics.
The object of the present work – the theoretical mechanics – also called the
general or rational mechanics, deals with the motion of the material systems comprised of
objects the deformations of which are negligible in practice.
The work represents the first edition of the Theoretical Mechanics course
presented by the author to the students of the Maritime University of Constanta as well as
support to the seminar classes, but it can be also consulted by industry and research
technical staff, students of higher education technical institutes or by industrial field high
schools teachers. The subject is presented in the usual sequence: statics, kinematics and
dynamics. Each of the 16 chapters is structured in four sections. The first one is devoted
to the theoretical notions related to the debated topics. It is followed by a set of
representative problems completely solved, structured in tight connection to the
theoretical solutions presented previously. The third section contains problems to be
solved, and in the end of the chapter, indications and solutions are given to these
applications. Thus, the reader is given the possibility to easily deepen knowledge and put
into practice the theorems and calculation methods of the mechanics. This structured
approach is also in keeping with the needs of other disciplines directly using mechanics
knowledge: Strength of Materials, Machine Elements, Theory of Mechanisms and
Measures, etc.
The author made his best efforts to follow the road of the “Romanian School of
Mechanics” characterized by a clear, systematic and mathematically rigorous
presentation, opened by Spiru Haret (1851-1912), Traian Lalescu (1882-1929) etc,
continued by Victor Valcovici (1885-1969), Caius Iacob (1912-1991) and a full series of
living mechanicists in full strength of creation, activating with great abnegation to the
progress of this science.
We hope that the present text shall not only succeed to be of real use to the
students preparing for the exams, but to awaken the interest of some or to maintain the
interest of others for throwing light upon the mystery and the charm of the Theoretical
Mechanics.
The author gives special thanks to Mrs. Engineer Elena Bogdan for the
competent desktop publishing and for putting her heart in publishing of this book.

The Author

3
4
.

CONTENTS

PREFACE ........................................................................................................................................3

CONTENTS .....................................................................................................................................5
1. INTRODUCTIVE NOTIONS ....................................................................................................9
1.1. Object of the Mechanics .........................................................................................................9
1.2. Principles of Classical Mechanics ........................................................................................10
1.3. Divisions of Classical Mechanics .........................................................................................10
1.4. Models of Classical Mechanics ............................................................................................10

2. VECTOR COMPUTING ELEMENTS ...................................................................................12


2.1. Scalars and Vectors...............................................................................................................12
2.2. Projection of a Vector onto a Plane or a Line .......................................................................14
2.3. Free Vectors Operations .......................................................................................................15
2.3.1. Free Vectors Addition ....................................................................................................15
2.3.2. Product of a Free Vector and a Scalar ............................................................................16
2.3.3. Scalar (inner) Product of Two Free Vectors ..................................................................17
2.3.4. Cross (outer) Product of Two Free Vectors ...................................................................19
2.3.5. Double Cross Product of Three Free Vectors ................................................................20
2.3.6. Mixed Product of Three Free Vectors ............................................................................20

PART I: STATICS

3. REDUCTION OF FORCE SYSTEMS ....................................................................................22


3.1. Forces and Systems of Forces ...............................................................................................22
3.2. Moment of a Force about a Point ..........................................................................................22
3.3. Moment of a Force about an Axis.........................................................................................24
3.4. Torsor of a Force about a Point.............................................................................................25
3.5. Torsor of a Force System about a Point ..............................................................................27
3.6. Minimal Torsor. Central Axis ...............................................................................................30
3.7. Cases of Reduction of a System of Arbitrary Forces ............................................................31
3.8. Reduction of Particular Systems of Forces ...........................................................................33
3.9. Solved Problems ...................................................................................................................36
3.10. Problems Proposed .............................................................................................................44
3.11. Indications and Answers .....................................................................................................46

4. CENTER OF MASS ..................................................................................................................49


4.1. Body Weight .........................................................................................................................49
4.2. Center of Mass (Weight) of a System of Material Points .....................................................49
4.3. Center of Mass of a Rigid Solid ............................................................................................50
4.4. Properties of the Center of Mass ...........................................................................................52
4.5. Static Moments. Theorem of the Static Moments .................................................................53
4.6. Solved Problems ...................................................................................................................54
4.7. Problems Proposed ...............................................................................................................58
4.8. Indications and answers ........................................................................................................59

5
5. EQUILIBRIUM OF THE MATERIAL POINT .....................................................................62
5.1. Introductive Notions ..............................................................................................................62
5.2. Equilibrium of the Free Material Point ..................................................................................62
5.3. Equilibrium of the Constrained Material Point. Principle of Connections. Classification of
Connections ..........................................................................................................................63
5.4. Equilibrium of the Material Point Subject to Frictionless Connections .................................64
5.5. Equilibrium of the Material Point Subject to Connections With Friction. Laws of Dry
Friction. Friction Cones. .......................................................................................................66
5.6. Solved Problems ...................................................................................................................69
5.7. Problems Proposed ...............................................................................................................72
5.8. Indications and answers ........................................................................................................74

6. EQUILIBRIUM OF THE RIGID ............................................................................................77


6.1. Equilibrium of the Free Rigid ................................................................................................77
6.2. Equilibrium of the Rigid Solid Subject to Constraints. Generals. .........................................78
6.3. Equilibrium of the Rigid Subject to Frictionless Constraints.
Simple Support. Joint. Embedment .......................................................................................78
6.4. Equilibrium of Rigid Subject to Friction Constraints.
Types of Friction. Equations of Equilibrium .........................................................................82
6.5. Solved Problems ...................................................................................................................89
6.6. Problems Proposed ...............................................................................................................95
6.7. Indications and Answers .......................................................................................................97

7. EQUILIBRIUM OF THE SYSTEMS OF MATERIAL POINTS AND OF SOLID


RIGIDS .....................................................................................................................................100
7.1. Conditions of Equilibrium of a System of Forces Acting Upon a System of Material Points
............................................................................................................................................100
7.2. Theorems and Methods for Solving Problems of Statics of Systems ..................................101
7.3. Solved Problems .................................................................................................................102
7.4. Problems Proposed .............................................................................................................110
7.5. Indications and Answers .....................................................................................................113

PART TWO: KINEMATICS

8. KINEMATICS OF THE ABSOLUTE MOTION OF THE MATERIAL POINT ............119


8.1. Trajectory, Velocity and Acceleration .................................................................................119
8.2. Analytical Expressions of Velocity and Acceleration in Different Systems of Coordinates121
8.3. Particular Motions of the Material Point.
Rectilinear Motion and Circular Motion..............................................................................125
8.4. Solved Problems .................................................................................................................128
8.5. Problems Proposed .............................................................................................................133
8.6. Indications and Answers ....................................................................................................134

9. KINEMATICS OF ABSOLUTE MOTION OF THE RIGID SOLID ................................137


9.1. General Motion of the Rigid ................................................................................................137
9.2. Particular motions of the rigid solid ....................................................................................141
9.2.1. Translational Motion of the Rigid..............................................................................141
9.2.2. Rotation of the Rigid..................................................................................................142
9.2.3. Helical Motion of the Rigid .......................................................................................146
9.2.4. Plane-parallel Motion of the Rigid ............................................................................148
9.2.5. Spherical Motion of the Rigid ...................................................................................151
6
9.3. General Motion of the Rigid ................................................................................................154
9.4. Solved Problems .................................................................................................................156
9.5. Problems Proposed .............................................................................................................168
9.6. Indications and Answers .....................................................................................................169

10. RELATIVE MOTION OF THE MATERIAL POINT ......................................................173


10.1. Definitions and Examples ..................................................................................................173
10.2. Relative Derivative and Absolute Derivative of a Vector .................................................174
10.3. Composition of Velocities .................................................................................................175
10.4. Composition of Accelerations ...........................................................................................175
10.5. Solved Problems ...............................................................................................................176
10.6. Problems Proposed ...........................................................................................................185
10.7. Indications and Answers ..................................................................................................187

11. RELATIVE MOTION OF THE RIGID SOLID ................................................................190


11.1. Generals .............................................................................................................................190
11.2. Composition of Velocities .................................................................................................191
11.3. Composition of Accelerations ...........................................................................................192
11.4. Particular Cases. Compositions of Instantaneous Motions ................................................193
11.5. Solved Problems ...............................................................................................................195
11.6. Problems Proposed ...........................................................................................................200
11.7. Indications and Answers ...................................................................................................201

PART THREE: DYNAMICS

12. DYNAMICS OF THE ABSOLUTE MOTION OF THE MATERIAL POINT ..............202


12.1. Dynamics of the Free Material Point .................................................................................202
12.2. Dynamics of the Motion of the Constrained Material Point ..............................................204
12.3. Solved Problems ...............................................................................................................204
12.4. Problems Proposed ...........................................................................................................209
12.5. Indications and Answers ...................................................................................................211

13. DYNAMICS OF THE RELATIVE MOTION OF THE MATERIAL POINT ...............214


13.1. Fundamental Equation of the Relative Motion ..................................................................214
13.2. Relative Rest ......................................................................................................................214
13.3. Inertial Systems of Reference ............................................................................................215
13.4. Solved Problems ...............................................................................................................216
13.5. Problems Proposed ...........................................................................................................221
13.6. Indications and Answers ...................................................................................................222

14. MECHANICAL MOMENTS OF INERTIA.......................................................................224


14.1. Definitions .........................................................................................................................224
14.2. Relationships Between Moments of Inertia .....................................................................225
14.3. Variation of the Moments of Inertia about Parallel Axes. STEINER’s Theorem..............225
14.4. Variation of the Moments of Inertia about Concurrent Axes.............................................227
14.5. Principal Axes and Moments of Inertia ............................................................................228
14.6. Solved Problems ...............................................................................................................229
14.7. Problems Proposed ...........................................................................................................234
14.8. Indications and Answers ..................................................................................................235

7
15. GENERAL THEOREMS IN THE DYNAMICS OF THE SYSTEMS OF MATERIAL
POINTS AND OF RIGID SOLIDS ............................................................................................237
15.1. Linear Momentum Theorem ..............................................................................................237
15.1.1. Linear Momentum of a System of Material Points and of a Rigid Solid ..............237
15.1.2. Linear Momentum Theorem. Theorem of the Motion of the Center of Mass ........238
15.1.3. Theorem of Linear Momentum Conservation ........................................................239
15.2. Theorem of the Angular Momentum .................................................................................240
15.2.1. Angular Momentum of a System of Material Points and of a Rigid Solid.
Koenig’s Theorem for the Angular Momentum .....................................................240
15.2.2. Expressions of the Angular Momentum in Different Particular Motions of the
Rigid ......................................................................................................................241
15.2.3. Theorem of the Angular Momentum in the Motion of a System of Material
Points or a Rigid related to a Fixed Reference .......................................................243
15.2.4. Theorem of the Angular Momentum in the Motion of a System of Material Points
or a Rigid around the Center of Mass ....................................................................244
15.2.5. Theorem of Angular Momentum Conservation ......................................................244
15.3. Work-Kinetic Energy Theorem .........................................................................................245
15.3.1. Kinetic Energy of a System of Material Points and of a Rigid Solid.
Koenig’s Theorem for Kinetic Energy....................................................................245
15.3.2. Elementary work and total work of a force.............................................................246
15.3.3. Forms of the Kinetic Energy in Various Particular Motions of the Rigid...............246
15.3.4. Work-Kinetic Energy Theorem in the Motion of a System of Material Points or
Rigid about a Fixed Reference…………………………………………………. 248
15.3.5. Work-Kinetic Energy Theorem in the Motion of a System of Material Points or
Rigid about the Center of Mass ..............................................................................250
15.3.6. Mechanical Energy Conservation ...........................................................................251
15.4. Solved Problems ................................................................................................................252
15.5. Problems Proposed ...........................................................................................................258
15.6. Indications and Answers ...................................................................................................260

16. DYNAMICS OF THE RIGID ..............................................................................................269


16.1. Dynamics of Rigid With Fixed Axis .................................................................................269
16.1.1. Problem Formulation ..............................................................................................269
16.1.2. Study of Rigid Motion ............................................................................................270
16.1.3. Determination of Reactions ....................................................................................270
16.1.4. Rotors Balancing ....................................................................................................272
16.2. Dynamics of the Rigid with a Fixed Point .........................................................................273
16.2.1. Problem Formulation ..............................................................................................273
16.2.2. Finding of the Equations of Motion. .......................................................................274
16.2.3. Determination of the Reaction ................................................................................276
16.2.4. Gyroscope: Stability, Gyroscopic Effect, Regular Precession Motion,
Gyroscopic Moment ...............................................................................................276
16.3. Dynamics of the Plane-Parallel Motion .............................................................................279
16.4. Rigid Dynamics in the General Motion .............................................................................280
16.5. Solved Problems ................................................................................................................281
16.6. Problems Proposed ............................................................................................................290
16.7. Indications and Answers ...................................................................................................291

BIBLIOGRAPHY ........................................................................................................................295

8
Theoretical mechanics – theory and applications

1. Introductive Notions
1.1. Object of the mechanics

Mechanics is a branch of the natural sciences that emerged from the earliest time.
This science played an important role in the development of the human civilization, its
birth and continuous development being a result of the practical problems related to
construction, transportation, etc occurred along the ages. Mechanics studies one of the
multiple features and manifestations of the matter, namely the motion, understanding by
this all the changes and processes happening into the universe, from the simple motion
through thinking. The matter is inexhaustible and so it is its knowledge process.
The classical theoretical mechanics or Newtonian Mechanics studies one of the
simplest forms of motion known as the mechanical motion, defined as the change of the
relative position of a body or a part thereof about another body considered as a reference
or in relation with a reference system. Development of mechanics took place
simultaneously with that of mathematics, the progress of each entailing the progress of the
others. From this point of view, the mechanics was a model for other basic natural sciences
(physics, biology, chemistry), they following an analogue process of mathematization.
Some of the greatest scientists of the world contributed the development of
mechanics. Their mere enumeration would require too much space. We will still recall the
scientists who had an essential contribution to the birth and development of this science.
Antiquity mentions Aristotle (384-322 BC) and Archimedes (287-212 BC). First
put forth the principle of inertia and the second is considered the founder of statics.
Renaissance period contains also big names: Leonardo da Vinci (1452-
1519), Nicolaus Copernicus (1473-1543), Johannes Kepler (1571-1630), CR
Huygens (1629-1695) etc.
Yet, Galileo Galilei (1564-1642) and Isaac Newton (1642-1727) remain the
founders of classical mechanics. Galileo established the basics of kinematics, velocity and
acceleration, studied falling bodies, motion on the inclined plane, formulated the principle
of inertia in its current form and curageously supported the Copernicus's heliocentric
concept. The one who extended study of mechanics to the whole planetary system and set
the first natural law written as a differential equation was, however, Newton. He defined
the fundamental principles of mechanics and discovered the universal law of gravity.
Reference scientists are also met and in the XVII-XIX centuries: Leonard
Euler (1707-1783), Jean LeRond D'Alembert (1717-1783), William Hamilton (1805-
1865), Joseph Louis Lagrange (1756-1813), Carl Friedrich Gauss (1777-1855), Henri
Poincare (1854-1912) etc.
Modern mechanics is related to the names of Max Planck (1858-1947), Erwin
Schrödinger (1887-1961) and Albert Einstein (1879-1955).
In our country, Spiru Haret (1851-1912), Andrei Ioachimescu (1868-
1943),Traian Lalescu (1882-1929), Victor Vâlcovici (1885-1970), Caius Iacob (1910-
1991), Radu Voinea distinguished in mechanics research.

9
Theoretical mechanics – theory and applications

1.2. Principles of classical mechanics

Classical mechanics is based on a number of principles (laws or axioms)


formulated by Isaac Newton in his fundamental work "Mathematical Principles of Natural
Philosophy" (1686). These are:
The principle of inertia (Newton’s 1st Law): "Every object continues in its state of
rest, or of uniform motion in a straight line, unless compelled to change that state by
external forces acted upon it.”
The principle of action force (Newton’s 2nd Law): "The change of motion is
proportional to the natural force impressed and is made in the direction of the straight
line in which the force is impressed"
Principle of Action and Reaction (Newton’s 3rd Law): "To every action there is
always opposed an equal and contrary reaction"
Parallelogram principle (Corollary I of Newton): “A body under the action of
two united forces describes the diagonal of a parallelogram in the same time that would
describe the sides under separate actions of forces”.

Notes: i) In the wording of the four principles, the body means a material point (see section
1.4).
ii) In the 2nd law mass is considered constant (as in all classical mechanics).
iii) The movement is relative to an absolute and motionless (fixed) reference system
iv) The second principle leads to the equation
→ →
ma =F (1.1)
→ →
where m is the mass of the material point, a its acceleration and F the exerted driving
force.
1.3 Divisions of classical mechanics

Methodologically, Newtonian mechanics is divided into three large chapters:

I) Statics – deals with the equilibrium of the material bodies, studying the systems
of forces reaching the equilibrium as well as reduction of the force systems.

II) Kinematics - study the motion of bodies without taking into account the forces
that act and their mass. Kinematics makes a geometric study of the mechanical
motion.

III) Dynamics - study the movement of bodies taking into account the forces that act
and their mass.

1.4. Models of classical mechanics

To simplify the study in mechanics, real bodies are substituted by their schematics,
called models. The results of the study on models extend on real bodies as they meet the
conditions considered in the study, and experience confirms the results. Classical
mechanics models are:

10
Theoretical mechanics – theory and applications

Material point - a point to which mass is assigned, representing the smallest


division of a body, and which retains the physical properties of the body while neglecting
its dimensions.
Material points system - a finite set of material points in interaction and
occupying a finite area of space.
Continuum - the model of a body for which it is admitted that any volume element
contains matter, that is it has mass.
Rigid body (rigid) - a continuous undeformable material, whatever system of
forces is acting on it.
The real solid body suffers deformation (elastic and plastic) but which classical
mechanics neglects. Using these models, the material point, the material points system and
the rigid body mechanics will be studied separately.
The class of the basic dimensions - length, time, mass – as well a series of
differentiated quantities such as force, speed, acceleration, momentum, work, kinetic
energy etc. are used in mechanics.

11
Theoretical mechanics – theory and applications

2. Vector computing elements


2.1. Scalars and vectors

Many physical quantities can be defined entirely by using a single number


invariable to the choice of a specific reference. Such quantities are called scalar
quantities or simply scalars. Representative examples of scalars used in mechanics are
mass, time, moment of a force about an axis, the mechanical energy, moments of inertia,
etc.
Many other physical quantities, however, present a geometric character, simply
specifying their numerical value is not sufficient for full characterization of such
quantity. It is the case of the vector quantities (or vectors). We can not conceive mechanics
discipline without discussing the concepts of force, moment of a force about a center,
speed, acceleration, momentum, angular momentum etc.

We shall consider that the classical mechanics working space is the three-
dimensional Euclidean space E 3 . To specify the position of a point M in space E 3 we
consider the following elements:
- A fixed point O of space E 3 , named origin ;
- A distance r from point O to point M;
- A sense of the segment [OM], namely from O to M.
Distance r, thus determined, is called the position vector of point M and is denoted
→ → → →
by r or OM . The absolute value (or magnitude) of the vector r is denoted by r . To

better characterize the vector r we use three non-coplanar lines, perpendicular at O two
by two, denoted by Ox, Oy and Oz, on which we consider a sense (positive sense). On the
positive side of axis Ox consider point A at a distance equal to unit from O. We denote the
→ →
position vector OA of this point by i and shall call it versor. Doing the same for axes Oy
→ →
and Oz we get the versors j and k (Figure T 2.1). Denote by M 1 , M 2 and M 3 the
projections of the point M on the axes Ox, Oy and Oz, respectively, and with x, y and z the
distances OM 1 , OM 2 and OM 3 (taken with the sign “+‘’ or ‘’-‘’ as the points M 1 , M 2 and
M 3 find on the positive or negative side of the axes Ox, Oy and Oz).

Since the vector OM 1 has the magnitude equal to the absolute value of the
→ → →
quantity x and the versor i has the magnitude 1 we shall agree to denote OM 1 = x i
→ →
thereby indicating that the vector OM 1 is determined by the scalar x and versor i .
→ → → →
Similarly, OM 2 = y j and OM 3 = z k .

For the position vector r of the point M we shall agree to write
→ → → →
r = x i + y j+ z k (2.1)

12
Theoretical mechanics – theory and applications

→ → → →
and will say that the vector r has the components x i , y j , z k and the point M the
coordinates x, y, z related to the axes of the three-dimensional reference frame Oxyz.

Representation (2.1) is called the analytical (or hyper complex) form of the vector r .
z
M3

N
M

v

M A

 r
k

 j M2 (d )
i
O y
M′ A′ N′

M1

Figure T 2.1 Figure T 2.2

From the position vector of the arbitrary point M we can step forward to the notion
of vector in general. Thus, considering the points M and N, we will define the vector
→ not →
MN = v as being the oriented segment MN. Projections of the points M and N onto the
→ → →
axes shall determine three vectors v x i , v y j , v z k , so we may write
→ → → →
v = vx i + v y j + vz k (2.2)

The scalar quantities v x , v y , v z are called coordinates of the vector v about the

axes of the reference Oxyz (or projections of the vector v ). The point M is called point of
→ →
application of the vector v and the point N is the terminal point of the vector v . Line MN

represents the line of action of the vector. If the point of application of vector v is
arbitrary, the vector is called free. If the application point is fixed, the vector is called tied.

Finally, if the point of application of vector v can be chosen anywhere on the line of
action, then the vector is called sliding.

The vector whose point of application coincides with its terminal point is called a
→ → → → →
null vector and denoted by 0 . If v = MN , then the vector NM , denoted - v , is called

opposite vector to the vector v .

13
Theoretical mechanics – theory and applications

2.2. Projection of a vector onto a plane or a line


→ →
The projection of a vector v = MN onto o line (d) or onto a plane (P) is obtained
by projecting its extremities onto that line or that plane. This provides new vectors, whose
sense will be determined by the sense of travel of the projection A’ of the point A on the

vector v , when A goes from M to N (Figure T 2.2).
→ →
Projection of a vector v onto an axis (d), denoted by prd v , is given by the
relation
→ →
prd v = v ⋅ cos θ (2.3)

where θ is the angle between the versor of the vector v and the versor of the axis (d). In
particular, we can write that
→ → → → → →
v x = prOx v = v cos α , v y = prOy v = v cos β , v z = prO z v = v cos γ (2.4)


where α , β , γ are the angles formed by the versor of the vector v with the versors
→ → →
i , j , k of the coordinate axes (Figure T 2.3).

Since cos 2 α + cos 2 β + cos 2 γ = 1 , we get that:

v 2 = v 2x + v 2y + v 2z (2.5)
z
C
vz D
 c
γ v

b B
β vy y
a
vx α
x A

y
x

Figure T 2.3 Figure T 2.4

→ →
Note i) If v = MN and M (x M , y M , z M ) , N (x N , y N , z N ) , then

→ → → →
v = (x N − x M ) i + ( y N − y M ) j + (z N − z M ) k (2.6)

14
Theoretical mechanics – theory and applications

2.3. Free vector operations

2.3.1. Free vectors addition


→ → → → → → → →
Definition 2.1: Let a = a x i + a y j + a z k and b = bx i + by j + bz k be two arbitrary

free vectors. The vector c given by the relation

( )
→ → → →
c = (a x + bx ) i + a y + by j + (a z + bz ) k (2.6)
→ → →
is called sum vector of the vectors a and b and the operation by which c is obtained from
→ →
a and b is called addition of vectors.
The addition operation is likely to have a geometric representation of a great
→ →
practical value. Vector c is the directed segment AC coinciding with the diagonal passing
→ → → →
through A of the the parallelogram constructed on vectors a = AB and b = AD as sides.
(Figure T 2.4). Construction thus obtained is called the parallelogram rule of addition.
→ → →
Notes : i) The sum vector c = a + b can be obtained otherwise. Thus, by moving vectors
→ → →
a and b parallel to their direction so that the point of application of vector b coincide
→ →
with the extremity of vector a we find that the sum vector c is the vector connecting the
→ →
point of application of vector a to the extremity of vector b . Summation rule thus
obtained is called the triangle rule (Figure T 2.5).
→ → →
ii) The triangle rule can be extended for a number of n vectors a 1 , a 2 ,..., a n , n ∈ N * , n ≥ 2 .
→ n →
The sum vector c = ∑a
i =1
i is the vector closing the polygonal contour consisting of the

vectors connected end to end in a random order, beginning with the point of application of
the first of them and ending with the terminal point of the last one. Thus, Figure T 2.6
→ → →
shows the polygon rule in the case of three vectors a 1 , a 2 , a 3 . Addition order is random.


a2

c c b a1 
b a3

   
a c = a1 + a 2 + a 3
a

Figure T 2.5 Figure T 2.6

iii) Addition of free vectors has the following properties:

15
Theoretical mechanics – theory and applications

→ → → → → → → → →
1) Is associative, i.e.  a + b  + c = a +  b + c , (∀) a , b , c ;
   
→ → → → → →
2) Is commutative, i.e. a + b = b + a , (∀) a , b ;
→ → → → → → →
3) Null vector, 0 , is a neutral element: a + 0 = 0 + a = a , (∀) a ;
→ → → → → → → →
4) (∀) a, (∃) − a (called opposite of a ) so that a + (− a ) = (− a ) + a = 0 .

We find, thus, that the addition operation determines a commutative group


structure onto the set of free vectors V. We shall write this group with (V, +). In this group,
→ → → → → →
the equation a + x = b has a unique solution which we write b − a = x and call it
→ →
difference between vectors a and b (Figure T 2.7).

α ⋅ a (λ > 0)
b
a
a
D

b a-b 
α ⋅ a (λ = 0)

A B 
a α ⋅ a (λ < 0)

Figure T 2.7 Figure T 2.8

2.3.2. Product of a free vector and a scalar


 → →
Definition 2.2: Let λ ∈ R and a ∈V . Application "⋅": R × V → V ,  λ , a  → λ ⋅ a is
 
called a scalar product.

Notes: iv) By λ ⋅ a we understand a free vector created as follows:
→ → →
1) if λ ≠ 0 and a ≠ 0 , then λ ⋅ a is a free vector having the same direction as vector
→ → →
a , same sense with a if λ > 0 , sense opposite to a if λ < 0 and the length

equal to λ ⋅ a (Figure T 2.8).
→ → → →
2) if λ = 0 or a = 0 , then λ ⋅ a = 0 .

v) Multiplication of free vectors by scalars has the following properties:


→ → → →
1) 1 ⋅ a = a ⋅ 1 = a , (∀) a ;

16
Theoretical mechanics – theory and applications

 → → →
2) α ⋅  β ⋅ a  = (α ⋅ β ) ⋅ a , (∀)α , β ∈ R, (∀) a ∈V ;
 
→ → → →
3) (α + β ) ⋅ a = α ⋅ a + β ⋅ a , (∀)α , β ∈ R, (∀) a ∈V ;
→ → → → → →
4) α ⋅  a + b  = α ⋅ a + α ⋅ b , (∀)α ∈ R, a , b ∈V .
 
→ →
Definition 2.3: Let a be a free vector. Vector u , defined by
→ 1 →
u = → ⋅a (2.8)
a

is called the versor of vector a .

→ →
Notes: vi) The versor u has the same direction and sense as vector a and the magnitude
equal to unit.

vii) A very useful way to project a free vector a onto the axes of a Cartesian reference
Oxyz , used in the following chapters, is given by the following algorithm (see also Figure
T 2.9) :

- let us consider a direction MN parallel to the direction of the vector a so that we
know the coordinates (x M , y M , z M ) and (x N , y N , z N ) of the points M and N ;
→ → →
→ 1 → (x N − x M ) i + ( y N − y M ) j + (z N − z M ) k
- u MN = ⋅ MN = ;

MN ( x N − x M )2 + ( y N − y M )2 + ( z N − z M )2
→ 1 → → → →
- since u MN = ⋅ a ⇔ a = a ⋅ u MN , then

a


a

 → → →

a= ⋅ ( x N − x M ) i + ( y N − y M ) j + ( z N − z M ) k 
(x N − x M ) + ( y N − y M ) + (z N − z M )
2 2 2  

2.3.3. Scalar (inner) product of two free vectors

→ →
Definition 2.4: Let us consider the vectors a and b . Angle ϕ ∈ [0, π ] determined
→ →
by two lines parallel to the directions of the vectors a and b , drawn through the same
→ →
point O, is called the angle of the vectors a and b (Figure T 2.10).

17
Theoretical mechanics – theory and applications

 
a
a 
z b
N(x N , y N , z N )

a
M (x M , y M , z M )

 O ϕ 
u MN b
y
O

Figure T 2.9 Figure T 2.10

→ → → →
Definition 2.5: Let us consider the free vectors a = a x i + a y j + a z k and
→ → → →
b = b x i + b y j + b z k . We shall call the scalar product of the two vectors the scalar
→ →
quantity written a ⋅ b and defined by the relation
→ →
a ⋅ b = a x bx + a y b y + a z bz (2.9)

Notes: viii) From the definition (2.9) the properties of the scalar product are as follows:
→ → → → → →
1) Commutativity : a ⋅ b = b ⋅ a , (∀) a , b ∈V ;

→ → →
 → → → → → → →
2) Distributivity over addition: a ⋅  b + c  = a ⋅ b + a ⋅ c , (∀) a , b , c ∈V .
 

ix) Starting from the definition (2.9) we can deduce the following geometric expression of
the scalar product:

→ → → →
→ →
a ⋅ b = a ⋅ b cos a , b  (2.10)
 

From formula (2.10) other properties of the scalar product are obtained:
→ → → → → → →
3) a , b ≠ 0 , a ⋅ b = 0 ⇒ a ⊥ b ;
→ → → → →
4) a ⋅ 0 = 0 ⋅ a = 0, (∀) a ∈V ;
5) Scalar product can be a positive, negative or zero scalar as the angle between the
π
two vectors is less than, greater than or equal to ;
2
2
→ → → →
6) a ⋅ a = a = a 2 , (∀) a ∈V ;

18
Theoretical mechanics – theory and applications

→ → →
7) With the versors i , j , k onto the axes of the Cartesian reference Oxyz the
following scalar products can be obtained:
→ → → → → → → → → → → →
i ⋅ i = j⋅ j = k ⋅ k = 1 , i ⋅ j = j⋅ k = k ⋅ i = 0 .

2.3.4. Cross (outer) product of two free vectors

→ → → →
Definition 2.6: Let us consider the free vectors a = a x i + a y j + a z k and
→ → → → → →
b = b x i + b y j + b z k . We shall call the cross product of the vector a by vector b , and
→ → →
shall write a × b as vector c whose projections on the axes of the Cartesian reference Oxyz
are:
c x = a y bz − a z b y , c y = a z bx − a x bz , c z = a x b y − a y bx (2.11)


Notes: x) Writing the vector c as
( ) ( )
→ → → →
c = a y b z − a z b y i + (a z b x − a x b z ) j + a x b y − a y b x k

we can notice that it is equal to the determinant

→ → →
i j k

c = ax ay az (2.12)
bx by bz

developed after the first line items.


xi) Using Definition 2.6 we deduce the following rule for determining the cross product of
→ →
vectors a and b :
→ → → → →
→ →
Cross product c = a × b is a vector of magnitude a ⋅ b sin  a , b  , normal to both
 
→ → →
of the factors and directed so that the trihedral  a , b , c  is right, that is, rotation of the
 

 → →
 →
vector a in the plane  a, b  with an angle not exceeding π in order to overlap b
 

coincide as sense with the sense of vector c (see Figure T 2.11).
xii) The cross product has the following properties:
→ → → → → →
1) Anticommutativity : a × b = − b × a , (∀) a , b ∈V ;

→ → → → → → → → →
2) Distributivity over addition : a ×  b + c  = a × b + a × c , (∀) a , b , c ∈V ;
 

19
Theoretical mechanics – theory and applications

→ →  → → →  → → →
3) λ ⋅  a × b  =  λ ⋅ a  × b = a ×  λ b , (∀)λ ∈ R, (∀) a , b ∈V ;
     
→ → → → → →
4) a × 0 = 0 × a = 0 , (∀) a ∈V ;
→ → → → →
5) a × b = 0 ⇔ (∃) λ ∈ R so that a = λ b ;
→ → → → →
6) If a ≠ λ b , then a × b is the area of the parallelogram built on the vectors a and

b (see Figure T 2.11).

2.3.5. Double cross product of three free vectors


→ → → → →
→ →
Definition 2.7: Let us consider the free vectors a , b , c . Vector w = a ×  b × c  is
 
called double cross product of the three vectors.

Notes: xiii) Using properties of the scalar product and cross product we can demonstrate
the following relation:

→ → → → → → → →
a×  b× c  = b ⋅  a ⋅ c  − c ⋅  a ⋅ b  (2.13)
     
→ → →
xiv) Vector w from Definition 2.7 is a vector coplanar with vectors b and c .

2.3.6. Mixed product of three free vectors

→ → →

→ → →

Definition 2.8: Given free vectors a , b , c , the scalar quantity a ⋅  b × c  is called
 
the mixed product of the three vectors.
Notes: xiv) Using (2.12) and the definition of the scalar product we get that :
ax a y az

→ → →

a ⋅  b × c  = bx b y bz (2.14)
 
cx c y cz

  
d = b×c
 
a×b
 A
a
B
 
e b ϕ C
ϕ
O  A  
a c b
o B

Figure T 2.11 Figure T 2.12

20
Theoretical mechanics – theory and applications

→ → →
xv) If a , b , c are non-coplanar vectors, then the magnitude of the mixed product represents
the volume of the parallelepiped that can be built on the three vectors considered with a
common point of application. (Figure T 2.12).

xvi) The mixed product has the following properties:



→ → →
 → → → → → → → → →
1) a ⋅  b × c  = c ⋅  a × b  = b ⋅  c × a , (∀) a , b , c ∈V ;
     

→ → →
 
→ → →
 → → →
2) a ⋅  b × c  = − a ⋅  c × b , (∀) a , b , c ∈V ;
   

→ → →
 →
 → →
 → → → → → →
3) λ a ⋅  b × c  = a ⋅  λ b × c  = a ⋅  b × λ c , (∀)λ ∈ R, (∀) a , b , c ∈V ;
     
 → →
  → →
 →
 → →
 →
→ → → → → →
4)  a 1 + a 2  ⋅  b × c  = a 1 ⋅  b × c  + a 2 ⋅  b × c , (∀) a 1 , a 2 , b , c ∈V ;
       

→ → →
 → → → → → →
5) a ⋅  b × c  = 0 ⇔ i) a = 0 or b = 0 or c = 0 ;
 
→ → →
ii) a , b , c are coplanar ;
→ →
iii) (∃)λ ∈ R so that b = λ c .

→ → →
 → → →
xvii) A notation used for the mixed product a ⋅  b × c  is  a , b , c  .
   

21
Theoretical mechanics – theory and applications

3. Reduction of force systems


3.1. Forces and systems of forces

The force is defined as a physical vector quantity that measures the mechanical
interaction between material bodies and that characterize the transmission of motion from
one body to another. Being a vector quantity, force has all the attributes of a vector: point
of application, direction, sense and magnitude.

The most encountered forces in nature are body weight, the universal attraction
force, friction forces, elastic forces, electromagnetic forces etc.
Forces that are exerted on a material system are divided into:
- external forces - if exercised by bodies outside the studied system on the bodies
in the system;
- internal forces - if exercised between the components of the same material
system.
In their turn, external forces are divided into directly applied forces (eg. gravity
force) and binding forces (due to geometrical nature restrictions imposed on bodies).
The totality of the forces acting on a material body is called a force system. Two
force systems are called equivalent if under the action of each of them, the body has the
same mechanical condition.
Any force acting on a rigid body has a character of a sliding vector, i.e. by sliding
it along its line of action, the force effect on the rigid remains the same.
To better characterize the effect of a force on a rigid it is necessary to introduce the
notions of moment of a force about a point and the moment of a force about an axis.

3.2. Moment of a force about a point (pole)

3.2.1. Definition

Definition 3.1: By definition, the moment of a force F about a point O is given by

the cross product between the position vector r of the point of application A of the force

and the vector F , namely
→ → → →
MO  F  = r × F (3.1)
 
 
→  →
The moment M O  F  is a tied vector, applied at the point O (Figure T 3.1). It is
 
 
→ →
normal to the plane determined by r and F , has the sense determined from right-hand-
screw rule and the magnitude:

22
Theoretical mechanics – theory and applications

M O  F  = r × F = r ⋅ F sin  r , F  = r ⋅ F sin α = F ⋅ d
→ → → → → → → → → → →
(3.2)
   

where distance d (from point O to the line of action of F ) is called the force arm.

( )
 
M0 F

O• F
A

O
d 
 F
r 
F
B
A α (d )

Figure T 3.1 Figure T 3.2


Denoting by Fx , Fy , Fz the projections of the force F onto the axes of the
Cartesian reference Oxyz and with x, y, z the coordinates of the point of application A of
the force, the analytical expression of the moment M O  F  is obtained:
→ →

 
→ → →
i j k
  z = ( y Fz − zFy ) i + (z Fx − x Fz ) j + (x Fy − y Fx ) k
→ → → → → → →
M O  F  = r× F = x y (3.3)
 
Fx Fy Fz

3.2.2. Properties of the moment of a force about a point

P1) Moment of a force about a point is zero if and only if the force line of action goes thru
that point (because force arm is zero).
P2) Moment of a force about a point does not change if the force moves its point of
application along its line of action.

Proof (see Figure T 3.2): Let A and B be two points on the line of action (d) of force F . If
the force is applied at A, then M O  F  = OA× F . For the force applied at B we have
→ → → →

 
M O  F  = OB× F =  OA+ AB  × F = OA× F + AB× F = OA× F ,
→ → → → → → → → → → → → →

   
→ →
because AB // F .

P3) Moment of a force about a point (pole) changes with the change of the pole.

23
Theoretical mechanics – theory and applications

Proof (see Figure T 3.3): Consider points O and O’. According to the Definition 3.1 we
may write that M O  F  = OA× F and
→ → → →

 
M O '  F  = O ' A× F =  O ' O + OA  × F = OA× F + O ' O× F .
→ → → → → → → → → → →

   
It follows that:
M O ' F  = M O  F  + O ' O× F
→ → → → → →
(3.4)
   

Moment of force F about the pole O’ remains unchanged only if O’O // (d).

3.3. Moment of a force about an axis



Definition 3.2: By definition, the moment of a force F about an axis ( ∆ ) is given

by the projection of the moment of the force F on the axis about a certain point O of the
axis, namely:
→ →
→ → → → → →  → → → → 
M ∆  F  = pr∆ M O  F  = M O  F  ⋅ u ∆ =  r × F  ⋅ u ∆ =  r , F , u ∆  (3.5)
         

where u ∆ is the versor of the axis ( ∆ ).

Notes: i) The definition given to the moment of a force about an axis makes sense because
assuming a point O’, different from O, located on the axis ( ∆ ) we get the same value
→
for the scalar M ∆  F  (see Figure T 3.4). Indeed:
 
 

 → → →
  → → → →  → → → → → → →
M ' ∆  F  =  r ' , F , u  =  O ' O + r , F , u  =  O ' O, F , u  +  r , F , u  = M ∆  F  ,
           
→ →
because O ' O // u .
ii) If we consider that the axis ( ∆ ) coincides with the axis Ox of the system Oxyz

and that the force F is applied at A(x, y, z), then

x y z
 →  → → →
M O x  F  =  r , F, i  = F x F y F z = y F z− z F y (3.6)
   
1 0 0
→ →
Similarly, M O y  F  = z F x − x F z , M O z  F  = x F y − y Fx , which means that the
   
components of the moment of a force about a point O on the axes of a triorthogonal
trihedral with the origin at O represents in fact the moments of the same force about the
axes of the trihedral.

24
Theoretical mechanics – Theory and applications

 
M Δ (F)
O′ (d )  
 
M′0 (F) O M 0 (F)
 
r F
  
F M′Δ (F) A

uΔ O′ 
O r′
A (Δ )
(d )

Figure T 3.3. Figure T 3.4

3.4. Torsor of a force about a point

3.4.1. Elementary operations of equivalence

Let’s consider a rigid solid under the action of an arbitrary system of forces. The
problem is to replace the force system for a simpler equivalent system, i.e. a system to
produce the same mechanical effect at every point in the rigid as the original force system.
To obtain equivalent force systems, but simpler, a series of operations are applied
to the forces that make up the system, so that the application of any of them results in an
equivalent system. These operations, called elementary operations of equivalence, are:
O1) The force acting on the rigid can be moved along its line of action.
O2) In the given system of forces two equal and directly opposite forces can be suppressed
or introduced.
O3) Several forces may be replaced by their resultant (obtained from the parallelogram
rule).

3.4.2. Couple of forces

Definition 3.3: A couple of forces is a system of two parallel forces, equal in


magnitude, acting on different lines of action.

Note iii): A couple of forces applied to a rigid tend to rotate it about an axis perpendicular
to the plane defined by the line of action of the two forces (Figure T 3.5).

Definition 3.4: The moment of a couple of forces means the sum of the moments
of the forces forming the couple about the same point.
We shall demonstrate the two useful properties below.

P1) Projection of a couple of forces on any axis is zero.



Proof: Let us consider an axis ( ∆ ) of versor u . We have that:
→ → →  → → →  → →
pr∆  F ,− F  = pr∆  F  + pr∆  − F  = F ⋅ u +  − F  ⋅ u = 0 .
       

25
Theoretical mechanics – Theory and applications

P2) The moment of a couple does not depend on the point the moment is calculated about
(it is a free vector, then).
Proof: Let us consider an arbitrary point O (see Figure T 3.5).
→ → → →
 →  → →  → → →
M O = OB× F + OA×  − F  =  OB − OA  × F = AB× F (independent of O).
   

Concluding, the moment M O of the couple is normal to the plane defined by the
lines of action of the forces, the sense is established with the right-hand-screw rule, and its
→ →
magnitude is M O = F ⋅ d , where d is the distance between the action lines of the forces

(called couple arm).



M0


F
A O
B

−F d

Figure T 3.5

3.4.3. Torsor of a force about a point: deduction and mechanical interpretation


Let us consider a rigid body (C) acted upon at point A by force F . We intend to
→ →
study the effect of force F at the point O. For that, we introduce at O the forces F and
→ → →
- F (based on equivalence operation no. 2). Force F from A and force - F from O are

→ → →
making up a couple characterized by its moment M O  F  = OA× F , perpendicular on the
 

plane determined by point O and force direction F (see Figure T 3.6).
(C)
z z
z  
M 0 (F)

~ 
   ~ F
F F F (π)
A

A A
O y O y O y
x  x
x −F

Figure T 3.6
26
Theoretical mechanics – Theory and applications

→ →
→
Thus, at O, the force F (equal to the one acting at A) and the moment M O  F 
 
shall act. The assembly of the two vectors forms what it is called the reduction torsor of

→ → → 
the force F applied at A, at point O. It is symbolically denoted by τ O  F  =  F , M O  or
   
→
F
τO  → → →
.
M O = OA× F


The reduction torsor represents the mechanical effect exerted by the force F
(acting at A) about point O.

3.5. Torsor of a force system about a point

3.5.1. Deduction

We shall consider a solid rigid acted upon at points A i , i = 1, n by the



forces F i , i = 1, n . We intend to determine the mechanical effect produced at a point O by
the simultaneous action of the n forces, i.e. the torsor τ O of the system (Figure T 3.7).


F1  
M1  F1
~ F2 ~
 
 M2 R
r1 
F2
• O′ 
 O 
r2 M0
A′n  O
A2
Mn 
Fn
O
rn

Fn

Figure T 3.7

For this we shall reduce all of the forces of the system one by one (see section
→ → → →
3.4.3.), and shall obtain at O the concurrent vectors F i , i = 1, n , and M i = OAi × F i , i = 1, n

respectively. Based on the equivalence operation no. 3, forces F i , i = 1, n , can be replaced
→ →
with their resultant R and moments M i , i = 1, n , can be replaced with the resulting

moment M O , thus obtaining the reduction torsor at point O of the force system

F i , i = 1, n :

27
Theoretical mechanics – Theory and applications

→ n →
R =
 i =1
∑Fi
τO  (3.7)
→ n →
M


O =
i =1
∑Mi

In conclusion, any system of forces acting on a rigid can be replaced by a single


force and a single vector moment applied at a point O conveniently chosen.

Note iv) If the rigid is also acted upon by p couples of force of moments M 'j , j = 1, p ,
then the reduction torsor has the following components:
→ n →
R =


i =1
Fi
τO  n → p
(3.8)
→ →


M O = ∑i =1
Mi + ∑M
j =1
'
j

3.5.2. Properties

P1) If the reduction pole changes from O to O’, then the second component of the torsor
changes as follows:
→ → →
 → →
 → →
n n n → n → →
M O' = ∑
i =1
O ' Ai × F i =∑  O ' O + OAi  × F i = O ' O×
i =1   i =1
Fi + ∑
i =1

OAi × F i =
→ → →
= M O + O ' O× R ,
so that:
→ n →
τ
 R = ∑ Fi
O'  i =1 (3.9)
→ → → →
M O ' = M O + O ' O× R

The resultant force is an invariant of the reduction operation of a force system at a


point (i.e. does not depend on the point of reduction).
→ →
P2) The scalar product of the resultant force R and resultant moment M O vectors is an
invariant of the reduction operation of a system of forces at a point.
→ → → → →
Proof: Performing the scalar multiplication of the relation M O ' = M O + O ' O× R by R
 → → →
and noting that  O ' O + R  ⋅ R = 0 , we get that:
 
→ → → →
M O ' ⋅ R = M O ⋅ R = constant (3.10)

28
Theoretical mechanics – Theory and applications

→ →
Product M O ⋅ R is called invariant trinomial and is the second invariant of the
→ →
reduction operation. To justify this denomination, let us consider the vectors R and M O
→ → → →
projected on the axes of the Cartesian reference Oxyz ( R = R x i + R y j + R z k ,
→ → → →
M O = M Ox i + M O y j + M O z k ) and use the relation (2.9). We get:

R x ⋅ M O x + R y ⋅ M O y + R z ⋅ M O z = constant (3.11)

P3) Projection of the resulting moment on the direction of the resulting force is an
invariant of the reduction operation.
→ →
Proof: Let us consider u R the versor of the direction of the resulting force R and M R the
projection of the resulting moment on the resultant (Figure T 3.8). Then:

→ →
→ → M O⋅ R
MR = M O⋅ u R = = constant.
R

Note v) M R is not an invariant independent of the resultant R and the invariant
→ →
trinomial M O ⋅ R , it resulting as a ratio of the two. For the operation of reducing a system of
→ → → →
forces at a point, we have two invariants R and M O ⋅ R or R and M R , namely.

 
M0 M0
(d )
  
uR MN R
 
MR  MR
O uR O

Figure T 3.8 Figure T 3.9

3.6. Minimal torsor. Central axis

3.6.1. Minimal torsor


If we perform the reduction of a force system F i , i = 1, n about different points of
the spaces it is found out that the reduction torsor is different due to the change of the
resulting moment (Figure T 3.9). The latter resolves into two components:

29
Theoretical mechanics – Theory and applications

→ →
M R - in the direction of resultant R ;
→ →
M N - in the direction found at the intersection of the plane normal at O on R with
→ →
the plane determined by the vectors R and M O .

We may write:
→ → →
MO =M R+ M N , M O = M 2R + M 2
N (3.12)

Since M R is an invariant of the reduction operation it follows that changes to the

resulting moment are due to the component M N which, depending on the point of
reduction, may take any position and any value in the plane normal to the resultant
→ →
direction. The minimum value of the resulting moment shall be obtained when M N = 0 :
→ →
M min = M R (3.13)

The reduction torsor formed by the resultant R and the minimum moment is called
minimal torsor (Figure T 3.10):

 →
 R

τ min :→ → → (3.14)
O⋅ R
→ → M →
M min = M R = M R ⋅ u R = ⋅R
 R 2

(d )

R
 
M min = M R
O

Figure T 3.10

3.6.2. Central axis

Definition 3.5: The geometric place of the points in space about which, by doing
the reduction, the minimal torsor is obtained, is called central axis.

Let us consider P(x, y, z) any point on the central axis relative to a Cartesian
reference Oxyz. Then:

30
Theoretical mechanics – Theory and applications

→ → →
i j k
→ → → → → → →
M P = M O + PO× R = M O x i + M O y j + M O z k+ − x −y −z
M Ox MOy M Oz

The resulting projections are:

(
M P x = M O x − y R z − zR y ) , M P y = M O y − (z R x − xR z ) ,
(
M P z = M O z − x R y − y Rx ) (3.15)

Noting that at a point of the central axis the components of the reduction torsor are
→ → →

collinear vectors  M P = M min // R  and considering the condition of collinearity of two
 
vectors, we get:
M Px M Py M Py
= = ⇔
Rx Ry Rz
M O x − y Rz + z R y M O y − z Rx + x Rz M Oz − x R y + y Rx
= = (3.16)
Rx Ry R z

that represents the equation of a line in space given by intersection of two planes.
→ →
Notes: vi) If R = 0 , then the notion of central axis becomes meaningless.
vii) The central axis can be also defined as the geometric place of points where the
→ → → →
resultant R and the resultant moment vector M O are collinear (if R⋅ M O ≠ 0 ).

3.7. Cases of reduction of a system of arbitrary forces

Reducing a system of forces acting on a rigid means to replace it with the simplest
system of forces having the same torsor as that of the given system. Since the resultant
moment appears like a couple of forces applied to a rigid, the following fundamental
theorem of reduction can be stated:

A force system applied to a rigid is equivalent to a single force, equal to the system
resultant, applied at an arbitrary point O and a couple of forces whose moment is the
resultant moment of the system about point O.

Depending on the magnitude of the two vectors, the following cases arise:

→ → → →
I) R = 0 , M O = 0 : A force system whose torsor is null is called equivalent to zero.
The forces of this system are in equilibrium and, therefore, a rigid acted on by such a
system of forces is in equilibrium.

31
Theoretical mechanics – Theory and applications

→ → → →
II) R = 0 , M O ≠ 0 : A force system whose reduction torsor only consists of the
→ →
moment M O , is equivalent to any couple acting on a plane (P) perpendicular to M O and
→ → →
whose moment coincides with M O as sense and magnitude, i.e. M O = F ⋅ d , where d is

the couple arm (Figure T 3.11).



M0

R

O F 
−F
O
d
(P )

Figure T 3.11 Figure T 3.12

→ → → → →
III) R ≠ 0 , M O = 0 : The reduction torsor only consists of the resultant R . The force

system is equivalent to a single force equal to R and applied at O (Figure T 3.12)
→ → → →
IV) R ≠ 0 , M O ≠ 0
→ → → →
a) R⋅ M O = 0 (i.e. R ⊥ M O ) : The force system is equivalent to a single force equal

to R , acting right on the central axis of the system (Figure T 3.13) because on its points,
the moment takes the minimum value which in this case is null
 → →

 R⋅ M O 
M
 min = M R = = 0 .
 R 
 
→ → → →
b) R⋅ M O ≠ 0 (i.e. R ⊥/ M O ) : To obtain an equivalent, but simpler system, let’s
→ →
resolve the vector M O into two components, i.e. a component M R on the direction of the

resultant and the component M N on plane (P) normal to the resultant direction (Figure T
→ →
3.14). The vectors M N and R , perpendicular to each other (case IV a), can be replaced for

the force R directed to the central axis so that the force system is equivalent to the minimal
→ →
torsor, i.e. by force R applied on the central axis and a couple formed from the forces F

32
Theoretical mechanics – Theory and applications


MR

and - F from plane (P), the couple arm being d = and the sense of forces chosen so

R


that couple moment equals to M R .
  
M0 R R

 M0 
R  F
~  MR ~
R d 
O •O  −F O′
O
O′ (P ) MN
(P )
axa centrală

Figure T 3.13 Figure T 3.14

3.8. Reduction of particular systems of forces

3.8.1. Reduction of concurrent forces


Definition 3.6: A set of n forces, F i , i = 1, n, whose lines of action are passing
through the same point O, make up a concurrent force system.
All of the forces can slide along their lines of action so that their application
points reach O. The moments of these forces about point O are null and the forces can be
replaced for their resultant. In conclusion, a concurrent force system is equivalent to a
→ n →
single force, equal to the resultant R = ∑F
i =1
i , whose line of action passes through point O
→ →
or is equivalent to zero if R = 0 .

3.8.2. Reduction of the coplanar forces


Definition 3.7: A set of n forces, F i , i = 1, n, whose lines of action are all
contained in the same plane (P), make up a coplanar force system (Figure T 3.15).

Denoting with Oxy the plane of the forces and remarking that:

→ → → → → →
r i = xi i + y i j , F i = Fi x i + Fi y j , i = 1, n (3.17)

we find for the torsor elements at O, the analytical expressions:

33
Theoretical mechanics – Theory and applications

→ n →  n  →  n  → → →
R =
 i =1
∑ ∑
 i =1   i =1

F i =  Fi x  ⋅ i +  Fi y  ⋅ j = R x i + R y j

τO  (3.18)
∑( )
→ n → → → →

n



M O = ∑
i =1
r i× Fi =
i =1
x i F i y − y i F i x ⋅ k = M O z k

→ →
Since R⋅ M O = 0 , then the coplanar force system can never be equivalent to the
minimal torsor. The reduction cases for the coplanar force system result from the cases of
reduction of systems any forces and are as follows:
→ → → →
I) R = 0 , M O = 0 : The force system is equivalent to zero.
→ → → → →
II) R = 0 , M O ≠ 0 : The force system is equivalent to a couple of moment M O .
→ → → →
III) R ≠ 0 , M O = 0 : The force system is equivalent to a single force, equal to the

resultant R , applied at O.
→ → → → →
IV) R ≠ 0 , M O ≠ 0 : The force system is equivalent to a single force, equal to R ,
applied to a point of the central axis.

The equations of the central axis result from particularization of the equations of
(
the general case R z = M O x = M O y = z = 0 . We get: )
x R y − y Rx = M O z
 (3.19)
z = 0
i. e. the line of equation x R y − y R x = M O z in the plane of forces.
 z
 A1
F1 
F1 A2 Fi
 (Δ )
A1 Ai
Ai F2
y  
z Fi  Fn
 u
An
ri 
Fn O y
A n (P )
O x x

Figure T 3.15 Figure T 3.16

34
Theoretical mechanics – Theory and applications

3.8.3. Reduction of the parallel forces



Definition 3.8: A set of n forces, F i , i = 1, n, whose lines of action are lines
parallel to each other form a parallel force system.
→ →
Let u be the versor of the common direction of the parallel forces F i , i = 1, n,
→ →
(Figure T 3.16). We may write that F i = Fi ⋅ u , i = 1., n , where Fi > 0 if the sense of force
→ →
F i coincides with the sense of the versor u and Fi < 0 if otherwise.
The elements of the reduction torsor at O are:
→ n →
 n  → →
R= ∑
i =1  i =1

F i =  F i  ⋅ u // u

(3.20)

→ n → → n → →  n →  → →
MO = ∑
i =1
r i× Fi = ∑
i =1  i =1 

r i × F i u =  F i r i  × u ⊥ u (3.21)

→ →
As a consequence, R⋅ M O = 0 and the reduction cases of a parallel force system
are the same as of the coplanar forces. In order to determine the central axis we shall use
the relation (3.21) and the observation that the resultant moment is null at a point of the
central axis. For an arbitrary point P of the central axis, we can write:
→ → → → →  n →  → →  n  → →
 i =1 

0 = M P = M O − OP× R ⇔  Fi r i  × u − r ×  Fi  ⋅ u = 0 ⇔
 i =1 

 n →  n  → → → n →  n  → →
    

∑ ∑
Fi r i −  Fi  ⋅ r × u = 0 ⇔
 i =1  
 ∑
Fi r i − 
 i =1 

Fi  ⋅ r = µ u ⇔
 i =1    i =1  
n →
→ ∑
i =1
Fi r i
µ

r = − ⋅u (3.22)
n n


i =1
Fi ∑
i =1
Fi

Denoting:
n →

µ
→ ∑
i =1
Fi r i
λ=− , rC=
n n


i =1
Fi ∑
i =1
Fi

we get the following equation, called the vector equation of the central axis:
→ → →
r = r C+λ ⋅ u (3.23)
Equation (3.23) represents the equation of a line passing through the fixed point C

35
Theoretical mechanics – Theory and applications


(called the center of parallel forces) and that is parallel to the axis (∆ ) whose versor is u
(see Figure T 3.17).
z

C λu

 r (Δ )
rc

u
O y

Figure T 3.17

In a Cartesian system Oxyz the coordinates of parallel forces center are:


n n n

∑i =1
Fi xi ∑
i =1
Fi y i ∑F i =1
i zi
xC = n
, yC = n
, zC = n
(3.24)
∑F
i =1
i ∑F
i =1
i ∑F
i =1
i

The center of the parallel forces has the following properties (without proof):
P1) Direction of all forces can be changed with the same angle and to the same sense, and
the central axis shall still pass through point C;
P2) The magnitude of all the forces can be multiplied by the same scalar and the center of
the parallel forces remains unchanged;
P3) Position of the center of the parallel forces does not depend on the selection of the
reference system (it is an intrinsic property of the force system).

3.9. Solved problems



R 3.1) A bent rod ABCD is acted upon by the force F , of magnitude 100 N. Determine the

moment of the force F about points A and B, respectively (Figure R 3.1). Dimensions are
in centimeters.
z

y 20 B
10

D
A
 8
D F E
z 4 10 
4 F
C
10

4
B O C y
8

x A
x

Figure R 3.1 Figure R 3.2


36
Theoretical mechanics – Theory and applications


Solution: As a first step, let us express the force F vectorially:

→ → → →
→ →
DE 8i + 4 j− 4k → → →
F = F ⋅ u DE = F ⋅ = 100 ⋅ = 81.6 i + 40.8 j − 40.8 k (N)
DE 8 2+ 4 2+ 4 2

As per definition 3.1, we have that:


→ → →
i j k

 →
 → →
M A F  = AD × F = 10 4
→ → →
− 8 = 163.2 i − 244.8 j + 81.6 k (N ⋅ cm )
 
81.6 40.8 − 40.8

More,
→ → → → → →
M F  = M A F  +BA × F
B
   
or
→ → →
i j k

 →
 → →
M B  F  = BD × F = 0 4
→ → →
− 8 = 163.2 i − 652.8 j − 325.6 k (N ⋅ cm )
 
81.6 40.8 − 40.8

R 3.2) On the diagonal AC of the square parallelepiped shown in Figure R 3.2 acts the

force F , of magnitude 1000 N. Find the moment of this force about the line going between
points D and C. Dimensions are given in centimeters.

Solution: Using definition 3.2, we get that:

→ → →  → → → → → → 
M DC  F  = prDC M C  F  = CA× F  ⋅ u DC = CA, F , u DC  (*)
       

But A(10, 0, 0), B(0, 20, 10), C(0, 20, 0) and D(10, 20, 10), so that:

→ → →
→ → → → − 10 i + 20 j + 10 k 1000  → → → 
CA = 10 i − 20 j , F = 1000 ⋅ = − i + 2 j + k  ,
100 + 400 + 100 6  
→ →
→ − 10 i − 10 k 1 → →
u DC = =− i+ k
100 + 100 2 

From (*) we now find that:

37
Theoretical mechanics – Theory and applications

10 − 20 0
→ 1000 2000 1000 10000
M DC  F  = − = (N ⋅ cm ).
  6 6 6 3
1 1
− 0 −
2 2
z
z
D G 
 
F4 M0
F3
−aP
E F


M O
C
 y 2P 2aP y
2P
F2

x R
A B

x F1

Figure R 3.3.1. Figure R 3.3.2.

R 3.3) On the cube OABCDEFG with the side a shown in Figure R 3.3.1., acts a system of
four forces, having the application points, directions and senses shown in the figure and the
magnitudes F 1= P, F 2 = P 2 , F 3 = P 3 , F 4 = 2 P as well as a couple of moment
M = a ⋅ P on direction OD. Requested:
a) Compute and represent the reduction torsor at point O;
b) What is the system of forces equivalent with?
c) Compute, if it is the case, the minimum moment;
d) Determine the equations of the central axis;
e) Determine the reduction torsor at point E.

Solution: a) The elements of the reduction torsor at point O are:


→ 4 →
R = ∑ F i
 i =1
τ O  4 → → 
→ →
 M O = ∑ M O F i  + M
 
i =1

Project in succession forces F i , i = 1,4 , on the axes of the system Oxyz and
determine then the moments of these forces about point O. Write the results in the table
T 3.1.

Force F 1 : Being directed on a parallel to the axis Oy, it is projected on this axis only.

Thus: F 1x = F1z = 0 , F1 y = P . Moment of force F 1 about point O is given by the relation:

38
Theoretical mechanics – Theory and applications

→ → →
i j k
→ →  → → →
M O F 1  = OA × F 1= a 0 0 =aP k
 
0 P 0

because A (a, 0, 0).


→ → →
→ → → AD −a i+ak → →
Force F 2 : F 2 = F 2 ⋅ u AD = F 2 ⋅ =P 2⋅ = −P i + P k , because
AD a2 + a2
→ → →
i j k
→ →  → → →
D(0,0,a). In addition, M O F 2  = OA × F 2 = a 0 0 =−aP j .
 
−P 0 P
→ → → →
→ → → DB a i + a j− a k → → →
Force F 3 : F 3 = F 3⋅ u DB = F 3⋅ =P 3⋅ = P i + P j− P k and
DB a2 + a2 + a2
→ → →
i j k
→ →  → → → →
M O F 3  = OD × F 3 = 0 0 a = − a P i + aP j .
 
P P −P

Force F 4 : Being parallel to the axis Ox, it is projected in true size on this axis, i.e.
→ → →
i j k
→ → → →  → → → →
F 4 = 2 P i . In addition, M O F 4  = OG × F 4 = 0 a a = 2 a P j − 2 aP k .
 
2P 0 0
→ → →
Moment M : M = a P k .
Force F x F y F z M x M y M z
→ 0 P 0 0 0 aP
F1
→ -P 0 P 0 -aP 0
F2
→ P P -P -aP aP 0
F3
→ 2P 0 0 0 2aP -2aP
F4
→ - - - 0 0 aP
M

∑ 2P 2P 0 -aP 2aP 0

Table T 3.1

39
Theoretical mechanics – Theory and applications

From table T 3.1 we deduce:


→ → →
 R = 2P i + 2 P j
τ O : → → →
 M = − aP i + 2 a P j
 O
→ →
and R = 2 P 2 , M O = aP 5 . The elements of the reduction torsor are represented in
Figure R 3.3.2.
→ → → → → →
b) Since R ≠ 0 , M O ≠ 0 and R ⋅ M O = 2 P ⋅ ( − aP) + 2 P ⋅ 2aP = 2aP 2 ≠ 0 , the force
system is equivalent to a minimal torsor.
→ → →
→ R⋅ M O R a P  → → 
c) M min = ⋅ =  i + j .
R R 2  
d) The general equations of the central axis (3.11) are rewritten for this problem as
follows:

− aP − y ⋅ 0 + z ⋅ 2 P 2aP − z ⋅ 2 P + x ⋅ 0 0 − x ⋅ 2 P + y ⋅ 2 P
= = ⇔
2P 2P 0
⇔ x − y = 0 ,4 z − 3a = 0 .

Note: The central axis was obtained as an intersection of the planes of equation x = y and
3a
z= .
4
e) The torsor at point E (a, 0, a) is composed of the resultant force vector
→ → →
R = 2 P i + 2 P j and the resultant moment vector:

→ → →
i j k
→ → → → → → → →
M E = M O + EO × R = − aP i + 2aP j + − a 0 − a = aP i − 2aP k .
2P 2P 0

R 3.4) The hexagonal plate OABCDE of side a in Figure R 3.4.1 is solicited by four forces
→ →
F i , i = 1,4 , situated in its plane and a couple of forces of moment M (perpendicular to the
3
plate plane). If F 1= F 3 = P , F2 = P 3 , F 4 = 2 P, M = a P, CF = λ a ( λ ∈ R) , it is
2
requested:

a) The reduction torsor at O (discussion depending on the values of parameter λ );


b) Central axis equation;
c) Graphical representation of the torsor at O and of the central axis, respectively.

40
Theorethical mecanics– theory and applications

y

M
A B

F1

F2 C
O x

F4
F

F3
E D

Figure R 3.4.1

Solution: a) Since it is a coplanar force system (located in the plane Oxy), the forces shall
have projections on the axes Ox and Oy only, and the moments of these forces about the
point O shall be normal to the plane Oxy, thus Fi z = M i x = M i y = 0, i = 1,4 . Magnitude of
the moments shall be computed with the relation (3.2) and the sense by the right-hand-
screw rule.

→ 1 3
Force F 1 : F 1x = F 1 cos 60 0 = P , F 1 y = F 1 sin 60 0 = P,
2 2
→ →  →
M O F 1  = 0 (O belongs to the line of action of force F 1 ).
 
→ 1 3
Force F 2 : F 2 x = F 2 cos 30 0 = P , F 2 y = F 2 sin 30 0 = P,
2 2
→ →  →
M O F 2  = 0 (O belongs to the line of action of force F 2 ).
 

Force F 3 : F 3 x = F 3 = P , F 3 y = 0 ,
→ →  →
M O F 3  = F 3 ⋅ d (O, ED) = F 3⋅ OE sin 60 0 = a P (having sense of axis
1
  2
Oz, the projection is positive) .

Force F 4 : F 4 x = 0 , F 4 y = − F 4 = −2 P ,
→ →  →  →  →
M O F 4  = F 4 ⋅ d O, F 4  = F 4 ⋅ CF sin 60 0 = 3 λ a P (since the sense is opposite to
   
the positive one on Oz, this projection is negative) .
→ → 3
Moment M : M = − M = − a P.
2
The results are summarized in Table T 3.4.

41
Theorethical mecanics– theory and applications

Forces Fx Fy Mz
→ 1
P 3
F1 P 0
2 2
→ 3
P 3
F2 P 0
2 2
→ P 0 3
F3 aP
2
→ - 2P 0 - 3λaP
F4
→ - - 3
M - aP
2
∑ P 3P - 3λaP

Table T 3.4

We get the torsor:


R→ → →
 = P i+ 3 P j
τ O → →
M O = − 3 λ a P k

b) Equation (3.14) of the central axis is in our case, of the form:


3 x− y =− 3 λ a.
c) Representation of the reduction torsor and of the central axis, depending on the values of
the real parameter λ is shown in the Figure R 3.4.2.

axa axa
y y y
centrala centrala

3 
 R 
R R
3λa
O O
O
−λa
x − λa x 3λa x
axa
centrala

λ=0 λ>0 λ<0

Figure R 3.4.2

42
Theorethical mecanics– theory and applications

→ ___
R 3.5) Given the parallel forces F i , i = 1,4 , of directions and senses as per Figure R 3.5,
magnitudes F1 = 4 P, F2 = 8P, F3 = 5P, F4 = 6 P and points of
 3a   3a 5a 
application A1 (3a,0,5a) , A 2  ,2a,0  , A3 ( 0,4a,0) , A 4  ,0,  respectively. Requested:
2  2 2

a) Determine the reduction torsor about point O;


b) Determine the parallel force center;
c) Determine vector equation of the central axis.

Solution: a) Forces project in true size onto Oz and give moments on the axes Ox
and Oy only. The projections of such vectors are given in the Table T 3.5.

Forces Fz Mx My
→ -4P 0 12 a P
F1
→ 8P 16 a P - 12 a P
F2
→ -5P - 20 a P 0
F3
→ 6P 0 -9aP
F4
∑ 5P -4aP -9aP

Table T 3.5

→ → → → →
τ O : R = 5P k , M O = −4 a P i − 9 a P j

b) Coordinates of the parallel force centre are:

4 4 4
∑ Fi x i ∑ Fi yi ∑ Fi zi
i =1 9 i =1 4 i =1
ξ= 4
= a ,η = 4
=− a,ζ = 4
= −a
5 5
∑ Fi ∑ Fi ∑ Fi
i =1 i =1 i =1
9 4 
It results that C  a , − a , − a  .
5 5 
c) The vector equation of the central axis is:
→ → → 9 → 4 → →
r = r C+ λ k = a i − a j + ( λ − a) k
5 5
where λ is an arbitrary constant.
The central axis is a line parallel to Oz that contains point C.

43
Theorethical mecanics– theory and applications

z z

C
 
F4 F3

A1 5a F2

 A4 
F2 O B
F1 A3
4a y
O  y
3a F3 A 
F1
A2 x

Figure R 3.5 Figure TC 3.1

3.10. Problems proposed

3.10.1. Conventional tests

TC 3.1) The triangular pyramid OABC in Figure TC 3.1, of sides AB = AC = a


 ∧  → → →
and masOAB  = 30 0 , is acted upon by the forces F 1, F 2 and F 3 oriented as shown in
 
→ 3 → →
the figure and of magnitudes F 1 = F 2 = F , F 3 = F . Requested:
2
a) Reduce the force system about point O and make a graphical representation of
the obtained reduction torsor elements;
b) What is the three force system equivalent with?
TC 3.2) Let us consider the rectangular parallelepiped OABCO’A’B’C’ of sides OA = 3a,
OC = 4a, OO’=12a solicited as in Figure TC 3.2, where
F1 = 10 P, F2 = 26 P , F3 = 21P and M= 6aP. On the side O’A’ acts a uniformly distributed
2P
load (parallel to Oz), changing in value from 0 at O’ to q = at point A’. Requested:
a
a) Reduction torsor at O;
b) What is the given system equivalent with?
c) What force and what moment must be introduced at O for the new system to be
reduced to a single resultant and the central axis to pass thru A and B?

z y F2
 2a
O′ M
O D A x
 C′
q z •
a a

A′ F1
B′  a
F2
C y
O 
 F4
 F3 B
F1 450
 C
A B F3 450
a

x
Figure TC 3.2 Figure TC 3.3

44
Theorethical mecanics– theory and applications

TC 3.3) A cantilever beam, having the dimensions in Figure TC 3.3, is solicited by the
→ ___
coplanar force system F i , i = 1,4 , having the directions and senses as in the figure and the
magnitudes F1 = P , F2 = 2 P , F3 = 3P , F4 = P 2 . Requested:
a) Reduction torsor at point O;
b) What is the force system equivalent with?
c) Central axis equation.

→ → → → → → → →
TC 3.4) Given parallel forces F 1= P i , F 2 = −5 P i , F 3 = 10 P i , F 4 = −2 P i , acting at
points A 1( a,2a,0), A 2 ( − a, a,2a), A 3( 2a, a,4a), A 4 ( 0,0,− a), respectively. Requested:
a) Compute and represent the reduction torsor at point O;
b) Determine the centre of the parallel forces and determine the central axis.

3.10.2. Multiple choice tests



TG 3.1) A system of arbitrary forces reduces to a single force equal to the resultant R that
→ → 
acts on the central axis of the system if the elements of the reduction torsor τ O R, M O 
 
about the arbitrary point O verifies the conditions:
→ → → → → → → → → → → → → → → →
a) R ≠ 0 , M O = 0 ; b) R ≠ 0 , M O ≠ 0 , R ⋅ M O = 0 ; c) R ≠ 0 , M O ≠ 0 , R ⋅ M O ≠ 0 ;
→ →
d) R // M O .

→ → →
TG 3.2) The point of application of the force F = 10 i + 3 j (N) has the position vector
→ → →
r = 5 j + 10 k (m) about the origin of the Cartesian system Oxyz. What is the moment of

the force F about the diagonal AD of the rectangular parallelepiped in Figure TG 3.2?
84 92 101
a) − ; b) 0 ; c) ; d) .
17 17 17
z
z
O′ C′
D
10 N

O
3m 5N C y
B′
A′
5N
10 N

y A
B
2m
x
x A 2m

Figure TG 3.2 Figure TG 3.3

45
Theorethical mecanics– theory and applications

TG 3.3) The sum of the moments of the force couples acting along the diagonals of the
rectangular parallelepiped in Figure TG 3.3 is:
→ → → → → →
a) M = 35.35 i + 22.36 j + 80.07 k ; b) M = 0 ;
→ → → → → → →
c) M = 50.14 i + 91.07 j − 14.65 k ; d) M = −23.14 i + 47.92 k .

3.11. Indications and answers

→ →
R = 0
TC 3.1) a) τ 0 → 3 → → → ;
M O = a F i + 3 j + k 
 4  

b) The force system is equivalent to a force couple of moment M O .

→ →
TC 3.2) a) The linearly distributed load is equivalent to the resultant force F = −3 P k ,
applied at the center D (2 a, 0, 12 a) of the distributed parallel forces.

Force Fx Fy Fz Mx My Mz
s
→ -6P 8P 0 0 0 24 a P
F1
→ 6P -8P 24 P 96 a P 0 - 24 a P
F2
→ 0 0 - 21 P 0 0 0
F3
→ 0 0 -3P 0 6aP 0
F
→ - - - 0 -6aP 0
M
∑ 0 0 0 96 a P 0 0

→ → → →
τ O : R = 0 , M O = 96 a P i

→ →
b) The force system is equivalent to a force couple of moment M O = 96 a P i .
→ → → → → → → →
c) Let us consider F ' = F '
x i+F '
y j+ F '
z k and M ' = M '
x i+M '
y j+ M '
z k the
force and the moment needed to be introduced at point O. The new reduction torsor at
O is:
τ O' : R ' = F 'x i + F 'y j + F 'z k , M O' = (96 a P + M 'x ) i + M 'y j + M 'z k
→ → → → → → → →

Because the resultant is parallel to the central axis (line AB), the conditions
F x = F 'z = 0 impose. The central axis equations are:
'

46
Theorethical mecanics– theory and applications

96 a P + M 'x + z F '
y M '
y M 'z − x F '
y
= '
=
0 F y
0
or
M '
M 'x + 96 a P
x= z
'
, z=− '
F y F y
→ →
In addition, R ' ⋅ M O=
'
F '
y ⋅M '
y. The central axis equations are: x = 3 a, z = 0.
Resulting:
M 'z = 3 a F '
y , M 'x = − 96 a P
→ →
The system is reduced to a single force if R ' ⋅ M O=
'
0 , i.e. M 'y = 0 (if F 'y = 0 , the
central axis notion becomes meaningless). Therefore
→ → → → →
F '= F '
y j , M ' = −96 a P i + 3 a F '
y k
'
where F y is an arbitrary constant.

TC 3.3) a)

Force Fx Fy Fz
s
→ 0 P aP
F1
→ 0 -2P -4aP
F2

F3
3P 0 3
2
(2 + 2 ) a P
→ -P -P -3aP
F4
∑ 2P -2P 3 ( 2
2
)
−1 a P

→ → → →  2 
τ O : R = 2 P i − 2 P j , M O = 3 − 1 a P
 2 
→ → → → → →
b) R ≠ 0 , M O ≠ 0 , R ⋅ M O = 0
The force system is equivalent to a single force, equal to the resultant, situated on the
central axis.
3
(
c) x + y = 2 − 2 a .
4
)
TC 3.4) a) See problem R 3.5.
→ → → → →
τ O : R = 4 P i , M O = 10 a P j − 7 a P k ;

47
Theorethical mecanics– theory and applications

 13 7 
b) The center of the parallel forces is the point C a, a, 8a  .
2 4 

TG 3.1) See “Cases of reduction of an arbitrary forces system”. Correct answer: b).

→ → → →
TG 3.2) The fact that M AD  F  = M A F  ⋅ u AD is used, where
   
→ → → → → →
M  F  = M O  F  + AO × F and A (2, 0, 0), D (0, 2, 3). Correct answer: c).
A
   

→ → → →
TG 3.3) M = M 1+ M 2 , where M 1 is a free vector, perpendicular to the plane ABC’O’, of

trigonometric sense and magnitude M 1= 5 N ⋅ 10 m = 50 N ⋅ m and M 2 is a free vector,
perpendicular to the plane BCO’A’, of trigonometric sense and magnitude
M 21= 10 N ⋅ 5 m = 50 N ⋅ m . Correct answer: a).

48
Theoretical mechanics – theory and applications

4. Centre of mass
4.1. Body weight

A body of mass m located at Earth's surface is subject to earth attraction. A


→ → →
force G = m g , called gravity, acts on it, where g is the terrestrial gravitational
acceleration. The value of the gravitational acceleration varies with latitude and altitude
but these variations are small and in calculations the average value g = 9,81 m / s 2 may be

considered. As concerns its direction, the vector g is directed about the center of the Earth,
with a small deviation due to rotation of the Earth.
If in a specific problem of mechanics bodies that occur are spread over a region

whose dimensions are negligible compared to the Earth, then the g vectors field can be
considered constant and, therefore, the weights of the different material points may be
considered as parallel forces and the results obtained in reduction of parallel force systems
may be used.

4.2. Center of mass (weight) of a system of material points



Let us consider the system of material points A i , i = 1, n , of weights G i , i = 1, n, and

the position vectors r i , i = 1, n (Figure T 4.1). Such a force system is equivalent to a
→  n  → →

unique force equal to G =  mi  ⋅ g = M ⋅ g (where M is the mass of the system), and
 i =1 
called gravity of the system. It acts at a point of the central axis that is a line parallel to the
direction of the forces and passes through the center of parallel forces of gravity defined
by:
n →

→ ∑G
i =1
i ri
rC = n
(4.1)
∑G i =1
i

Point C is called in this case centre of gravity of the system of material points.
Substituting Gi = m i ⋅ g , i = 1, n, in (4.1) and simplifying by g we get:
n →

→ ∑
i =1
mi ri
rC = n
(4.2)
∑m i =1
i

49
Theoretical mechanics – theory and applications

z z Axa
m1
mt centrala
 mn C
G1 
rc  
   G = M⋅g
ri Gt Gn

O y
≅ O
y

x x

Figure T 4.1

By projecting relation (4.2) onto the axes of the Cartesian system Oxyz we obtain
the coordinates of the point C:
n n n

∑i =1
mi xi ∑
i =1
mi yi ∑m
i =1
i zi
xC = n
, yC = n
, zC = n
(4.3)
∑m i =1
i ∑m i =1
i ∑m i =1
i

Expressions (4.2) and (4.3) do not depend on g, but on the distribution of the
masses m i , i = 1, n only. We can enter now the notion of center of mass of a system of
material points, point C is therefore the mass center of the system of masses m i , i = 1, n .
The concept of center of mass is more general than the center of gravity, which
makes sense only for the masses in gravitational field (not necessarily terrestrial). It is
inappropriate to talk about the center of gravity of the planetary system but we can talk
about its center of mass (which is very close to the center of mass of the Sun). Center of
mass and center of gravity coincide, still, for a system of material points in uniform
gravitational field.

4.3. Center of mass of a rigid solid

In case of rigid solid (Figure T 4.2) formulas (4.2) and (4.3) change in the sense
that the sums become integrals:


rC =
∫ ( D)
r dm
(4.4)
∫ ( D)
dm

xC =
∫( D)
x dm
, yC =
∫ ( D)
y dm
, zC =

( D)
z dm
(4.5)
∫( D)
dm ∫ ( D)
dm ∫
( D)
dm
where (D) is the domain occupied by the rigid solid.

50
Theoretical mechanics – theory and applications

To study the mass center of rigid solid is necessary to introduce a new quantity,
namely density. In terms of density, bodies are divided into homogeneous bodies (constant
density) and inhomogeneous bodies (variable density).
The following categories of densities can be defined:
- for bodies with weight spatially distributed on a volume V (all three
dimensions have the same order of magnitude), volumetric density:
dm  M
ρ V=  sau ρ V =  (4.6)
dV  V 
- for bodies with weight spread over a surface of area A (one of the sizes is
negligible compared to the other two), surface density:
dm  M
ρ S=  sau ρ S =  (4.7)
dA  A
- for bodies with weight distributed on a curve of length L (two dimensions are
negligible compared with the third) linear density:
dm  M
ρ L=  sau ρ L =  (4.8)
dL  L
The definitions in brackets correspond to homogeneous bodies.

Substituting into (3.4) the mass element dm, with:

dm = ρ V dV , dm = ρ S dA , dm = ρ L dL (4.9)

∫ (D )
dV = V , ∫ ( D)
dA = A , ∫
( D)
dL = L (4.10)

we obtain, after simplifying the densities, the relations:

→ → →

rC =

( D)
r dV
,

rC =

( D)
r dA
,

rC =

( D)
r dL
(4.11)
V A L
valid for homogeneous bodies with spatially, superficially or linearly distributed mass. In
the case of inhomogeneous bodies the relation (4.4) is used, where dm is obtained from
(4.9).
z

dm

dG
 C
r 
rc
O  y
G

Figure T 4.2

51
Theoretical mechanics – theory and applications

4.4. Properties of the center of mass

Center of mass of a material point system or a rigid solid has the following
properties (without proof):
P1) The position of center of mass does not depend on the chosen reference
system, being an intrinsic point of the material system.
P2) If the body (material points system) has a plane, an axis or center of symmetry,
then the center of mass lies in that plane, on that axis or at that point.
P3) If a system (S) of material points is composed of n
subsystems (S1 ), (S 2 ),..., (S n ) , whose masses M i , i = 1, n and mass centers Ci , i = 1, n are
known, then the center of mass of the system (S) can be obtained considering that the
masses of the component systems would concentrate at the centers of mass (Figure T 4.3),
i.e.:
n →

→ ∑M
i =1
i ⋅ r Ci
rC = n
(4.12)
∑M i =1
i


where r C i are the position vectors of the points C i , i = 1, n .
P4) If a system of material points (S) may be considered as originating from a
system of material points (S 1 ) from which another system (S 2 ) was removed and if the
masses M 1 and M 2 , and the centers of mass C 1 and C 2 thereof are known (T Figure
4.4), then the center of mass of the system (S) can be obtained considering that the masses
M 1 and − M 2 are concentrated at C 1 and C 2 and:
→ →
→ M r C − M2 r C2
rC = 1 1 (4.13)
M1 − M 2

(S1 )
z
• C S1 (m1 )...... Si (m i )
1 z C (S2 )
•1
• Ci C
• 2

rC
 1
rC i
Sn ( m n )

•C • Cn rC
 O 2
rC y
y

x
x

Figure T 4.3 Figure T 4.4

52
Theoretical mechanics – theory and applications

4.5. Static moments. Theorem of the static moments

In relation (4.3) giving the coordinates of the center of mass of a system of


n n n
material points we have the sums ∑
i =1
mixi , ∑
i =1
miyi , ∑m
i =1
izi of points masses and their

distances to their coordinate planes. These sums are called static moments about
coordinate planes and are denoted as follows:
n n n
SO x y = ∑
i =1
m i z i , SO y z = ∑
i =1
m i x i , S O zx = ∑m
i =1
i yi (4.14)

→ n →
In addition, the sum S O = ∑m
i =1
i r i in (4.2) is called static polar moment.

Static moments characterize the distribution of mass of a system of material points.


From (4.2) and (4.3) we get that:
→ →
SO = M ⋅ rC , SO x y = M ⋅ zC , S O y z = M ⋅ xC , S O z x = M ⋅ yC (4.15)
i.e.:
i) Static moment of a system of material points about point O equals to the
mass of the system times the position vector of the center of mass about
point O
ii) Static moment of a system of material points about a plane equals to the
product between the mass of the system and the distance from its center of
mass to that plan.

Statements i) and ii) represent the theorem of the static moments

Notes: 1) Those above will also apply to a rigid solid (except that sums become integrals)
n
2) C ∈ Oyz ⇔ ∑m
i =1
i xi = 0 (4.16)

Similar relations take place if C ∈ Ozx and C ∈ Oyx .

4.6. Solved problems

R 4.1) Let the plane system of material points in Figure R 4.1 be, of
___
masses mi = i ⋅ m , i = 1,4 , located at points of the circle of equation x 2 + y 2 = 7a 2 and of
the hyperbola of equation x 2 − y 2 = a 2 . Determine the coordinates of the center of mass
of this system.
Solution: The coordinates of the points at which material points are found are
obtained by solving the system composed of the equations of the circle and of the
hyperbola. We get:

53
Theoretical mechanics – theory and applications

x 2 + y 2 = 7 a 2

 ( ) (
⇒ A 1 − 2a, 3 a , A 2 − 2a ,− 3 a , A 3 2a, 3 a , A 4 2a,− 3 a ) ( ) ( )
x 2 − y 2 = a 2

The results are synthesized in table T 4.1.

Point mi xi yi m i xi m i yi
1 M -2a 3a -2ma 3 ma
2 2m -2a - 3a -4ma - 2 3 ma
3 3m 2a 3a 6ma 3 3 ma
4 4m 2a - 3a 8ma - 4 3 ma
∑ 10 m 8ma - 2 3 ma

Table T 4.1
We find thus:
4 4
∑ mi x i ∑ m i yi
4 3 4 3 
a ⇒ C a ,−
i =1 i =1
xC = = a , yC = =− a .
4 5 4 5 5 5 
∑ mi ∑ mi
i =1 i =1 z

60

y
O
y

•m •
A1 A2
80

m2
1
x B
C 30
100

•A m •
O1 D
3
m4 x 20

3 A4
E

40

Figure R 4.1 Figure R 4.2

R 4.2) Let the homogenous beam be of dimensions in Figure R 4.2, given in cm.
Determine the coordinates of the center of mass.
Solution: The homogenous beam may be considered as being composed of:
- line segment OA, 60 cm in length, situated on the axis Oz ;
- line segment OB, 80 cm in length, situated on the axis Ox ;
- line segment BC, 100 cm in length, situated in the plane Oxy ( BC // Oy ) ;
- semicircle CD, 30 cm in radius, situated in the plane Oxy ( CD // Oy ) :
- line segment DE, 20 cm in length, situated in the plane Oxy ( DE // Ox ) ;
- line segment EF, 40 cm in length, situated on a parallel to Oz.

54
Theoretical mechanics – theory and applications

The six components will be numbered from 1 to 6. Coordinates of the center of mass are
calculated with the formulas:
6 6 6
∑l i x i ∑l i y i ∑l i z i
i =1 i =1 i =1
x C= 6
, yC = 6
, zC = 6
∑l i ∑l i ∑l i
i =1 i =1 i =1
where x i , y i , z i represent the coordinates of center of mass for bar no. i and l i its length .
Write the data in Table T 4.2.

Body li xi yi zi li xi li yi l i zi
1 60 0 0 30 0 0 1800
2 80 40 0 0 3200 0 0
3 100 80 50 0 8000 5000 0
4 30 π 60 130 0 2400 π -1800 3900 π 0
80 -
π
5 20 90 160 0 1800 3200 0
6 40 100 160 - 20 4000 6400 - 800
∑ 30(10 + π ) 800 (19 + 3 π ) 100 (146 + 39 π ) 1000

Table T 4.2
Resulting values:

80 (19 + 3π ) 10 (146 + 39π ) 100


xC = , yC = , zC =
3 (10 + π ) 3 (10 + π ) 3 (10 + π )

R 4.3) Determine position of the center of mass for the homogenous plate in Figure R4.3.
Dimensions are in cm.
Solution: Determine the center of mass using the formulas:
4 4
∑ Ai xi ∑ Ai y i
i =1 i =1
xC = 4
, yC = 4
∑ Ai ∑ Ai
i =1 i =1

where x i , y i represent the coordinates of center of mass for plate no. i and Ai its area.
The plate in Figure R 4.3 was divided into four basic plates (see table below).

135 π
2333
3 − 108π 1603
3 + 2
x C= 9π
≅ 4.024 , y C= 9π
≅ 6.852
116 − 4 116 − 4

55
Theoretical mechanics – theory and applications

Body Ai xi yi Ai xi A i yi

1 120 6 5 720 600

27 π 4 4 135π
− 10 + 9+
2 4 3π 3π -9 2
4 2 16 8
3 -4 - -
3 3 3 3

8 8
-9 π 12 − 72 − 108π
4 π π - 72

9π 2333 1603 135π


116 − − 108π +
∑ 4 3 3 2

Table R 4.3

Note: Figures not crosshatched correspond to the bodies artificially introduced as a


completion to a basic figure (here, a rectangle). Their areas shall be considered with the
sign “- “.
y
r1 = 3 y a)
x+
/2(
: y =d
D

10 x
r2 = 6 O
2 x
P : y = a2 − x2
4
12

Figure R 4.3 Figure R 4.4

R 4.4) Determine position of the center of mass for the homogenous flat plate in Figure
a
R4.4, bounded by the parabola of equation y = a 2 − x 2 and the line y = ( x + a) , where
2
a> 0 is a given constant.
x2 x2
∫ x (g ( x) − f ( x) ) dx 1
2 ∫ (g 2 ( x) − f 2 ( x) dx)
x1 x1
Solution: x C= x2
, y C= x2
,
∫ (g ( x) − f ( x) ) dx ∫ (g ( x) − f ( x) ) dx
x1 x1

56
Theoretical mechanics – theory and applications

a
where g ( x) = a 2 − x 2 , f ( x) = ( x + a) and x 1 and x 2 , respectively are the solution of the
2
a
equation g (x) = f (x) , that is x 1= − a , x 2 = . Following the computation of the integrals it
2
a 23 2
is found that x C = − , y C = a .
2 30
R 4.5) Determine position of the center of mass for the semispherical homogenous body of
radius R in Figure R 4.5.
Solution: Axis Oz being axis of symmetry we may write that
∫D z dV
x C= y C= 0, z C= , where D is the domain occupied by the body. Using the
∫D dV
spherical coordinates system ( ρ , θ , ϕ ) and remarking that:

z = ρ sin ϕ , dV = ρ 2 cos ϕ dρ dθ dϕ
π
0 ≤ ρ ≤ R , 0 ≤ θ ≤ 2π , 0≤ϕ ≤
2
We find that:
π
R 2π 2
π R4
∫D ∫
z dV = ρ 3 dρ ⋅ ∫ ∫
dθ ⋅ sin ϕ cosϕ dϕ =
4
0 0 0
π
R 2π 2
2π R 3
∫D ∫
dV = ρ 2 dρ ⋅ ∫ ∫
dθ ⋅ cosϕ dϕ =
3
0 0 0
3
so that z = R.
8 y
z

4r
2r
ρ
O O1 3r x
ϕ 5r
O y

Figure R 4.5 Figure TC 4.1

57
Theoretical mechanics – theory and applications

4.7. Problems proposed

4.7.1 Conventional tests

TC 4.1) Determine position of the center of mass for the homogenous beam in Figure TC
4.1.

TC 4.2) Determine position of the center of mass for the homogenous flat plate in Figure
TC 4.2.
z r1
y C
4a

60 0
O2
r2
6a y
O
B

a
2a

a x
4a A
x

Figure TC 4.2 Figure TC 4.3

TC 4.3) Let the body in Figure TC 4.3 be, composed of three homogenous flat plates made
from dissimilar materials. The plates in the planes Oxy and Oyz have the superficial
density ρ and the plate in the plane Oxz has the superficial density 2 ρ . Knowing the
radiuses r 1= r 2 = l , determine the coordinates of the center of mass.

TC 4.4) Determine position of the center of mass for the homogenous flat plate in Figure
TC4.4.
y
x
2 2 3a
x y
+ =1
4a 2 9a 2
r ( y) y
O

2a a 3a z y
O x
1

a ⋅ y = x 2 − xa − 6a 2

Figure TC 4.4 Figure TC 4.5

58
Theoretical mechanics – theory and applications

TC 4.5) Determine position of the center of mass for the homogenous body of revolution
in Figure TC 4.5 knowing that the law of variation of the radius (distance from the axis Oy
to the body surface) is r ( y) =
1 2
y (m).
20

4.7.2. Multiple-choice tests


TG 4.1) Determine the abscissa of center of mass for a straight beam AB of length L,
where the density varies linearly from value ρ 1 at end A to value ρ 2 at end B (Figure TG
4.1).
(ρ + 2 ρ 2 )L ; b) x = (2 ρ 1+ ρ 2 )L ; c) x =  ρ 1 + ρ 2  L ;
a) x C = 1
3 (ρ 1+ ρ 2 )
C
3 (ρ 1+ ρ 2 )
C  
 ρ 2 ρ1 
ρ 1+ ρ 2 L
d) x C = .
2 ρ1ρ 2 2 y
x = R ⋅ cos θ
y
y dl = R ⋅ dθ

l α dθ
θ 2a
x dx O •
α x
⋅ z C

A dm B x R 4a 2a x
z

Figure TG 4.1 Figure TG 4.2 Figure TG 4.3

TG 4.2) Determine the abscissa of center of mass for a homogenous beam AB, of constant
section, circle arc shaped, of radius R and central angle 2 α , angle α being expressed in
radians (Figure TG 4.2).
2R 2 sin α sin α
a) x C = 0 ; b) x C = ; c) x C = R ; d) x C = R .
3 3 α α
TG 4.3) The center of mass of the homogenous flat plate, having the shape and dimensions
in Figure TG 4.3 has the coordinates:
2a 20 a 5a 5a 4a
a) x C = y C = ; b) x C = , yC = ; c) x C = , yC = ; d)
3 13 13 13 13
38 a 14a
xC= , yC = .
15 15

4.8. Indications and answers

TC 4.1) Divide the beam into four semicircles of radiuses 2r, 3r, 4r and 5r and proceed as
4r 2r
when solving problem R 4.2. It is obtained that x C = , y C= − .
7 π

59
Theoretical mechanics – theory and applications

TC 4.2) Complete plate to a full rectangle. The components are: a rectangle of sides 6a
and 4a, a right triangle of legs 2a and a, a sector of a circle of radius 4a and angle at the
center 60 0 and a sector of a circle of radius a and the angle at the center 270 0 (see also
solution for the problem R 4.3). Areas of the triangle and of the sector of circle of radius
4a are considered as negative. Coordinates obtained:

(
4 144 + 9 π − 32 3 a ) 4 (245 − 48 π ) a
x C= , y C= .
23 (12 − π ) 23 (12 − π )

TC 4.3) To determine the coordinates of the center of mass, the following formulas are
used:

4 4 4
∑ ρ i Ai x i ∑ ρ i Ai y i ∑ ρ i Ai z i
i =1 i =1 i =1
x C= 4
, y C= 4
, zC = 4
∑ ρ i Ai ∑ ρ i Ai ∑ ρ i Ai
i =1 i =1 i =1

as for a plate type homogeneous body the following relation takes place: m i = ρ i A i .
It is considered that the flat plate in the plane Oxz was obtained from a square
plate of side l (plate 1) from which a quarter of a disk of radius r 2 (plate 2) was removed.
Area of the plate 2 is considered with a negative sign. Following values are obtained:
2(11 − 3π ) 2 2(4 − π )
x C= l ≅ 0,153 l , y C = l ≅ 0,292 l , z C = l ≅ 0,250 l
3(10 − π ) 10 − π 10 − π

TC 4.4) The plate is composed from four parts: a right isosceles triangle of legs 3a in
quadrant I, internal of the quarter of ellipse in quadrant II, internal of the arc of a parabola
in quadrants III and IV and the disk of radius a with center in origin (whose area is
considered negative (see R 4.3 and R 4.4 solving). Following coordinates are obtained:
131 a 499 a
x C= y C= −
2 (3 π + 152) 2 (3π + 152)
, .

TC 4.5) Divide the body into infinitesimal cylinders of radius r and height dy by sections
∫D y dV 3
with planes parallel to plane Oxz. We have that: x C = z C = 0 , y C = = , since:
∫D dV 4
π π
∫D dV = ∫0 π r 2 dy =
1
60
, ∫D y dV = ∫0 π r 2 y dy =
1
80
,

D, being the domain in the space occupied by the revolution body.

60
Theoretical mechanics – theory and applications

TG 4.1) Divide the beam into infinitesimal cylinders of length dx. We have
∫AB x dm ( ρ 1+ 2 ρ 2 ) L
that y C = z C = 0 , x C = =
∫AB dm 3 ( ρ 1+ ρ 2 )
, since:

L  ρ 2− ρ 1  ( ρ 1+ 2 ρ 2 ) L 2
∫AB x dm = ∫AB x ρ ( x) dx = ∫0 x 
 L
+ ρ 1  dx =
 6
L  ρ 2− ρ 1  ( ρ 1+ ρ 2 ) L
∫AB dm = ∫AB ρ ( x) dx = ∫0 
 L
+ ρ 1  dx =
 2
.

Correct answer: a).

TG 4.2) Divide the beam into infinitesimal arc elements of length dl = R dθ . It is obtained
∫AB x dl sin α
that: y C = z C = 0, x C = =R , since:
∫AB dl α
α
∫AB dl = ∫−α R dθ = 2 α R , ∫AB x dl = ∫−αα R 2 cos θ dθ = 2 R 2 sin α .
Correct answer: d).

TG 4.3) Divide the plate in two: rectangle of sides 2a and 4a and the right and isosceles of
legs 2a. Correct answer: d)

61
Theoretical mechanics – theory and applications

5. Equilibrium of the material point


5.1. Introductive Notions

We shall define below some of the terms used in this chapter.


A free material point is the material point that can take any position in space.
Material point subject to connections (constrained) is a point on which a
geometric restriction is imposed (such as an obligation to remain on a surface or a curve).
The number of degrees of freedom is the number of independent scalar
parameters necessary to determine, at a given instant of time, the position of the material
point (or rigid).

The position of a material point in space is determined by means of three


independent scalar parameters such as Cartesian coordinates x, y, z. Consequently, a free
material point in space has three degrees of freedom.
The position of a material point on a surface is given by two independent scalar
parameters. Thus, if the surface is the plane Oxy it is sufficient to know the coordinates x
and y of the position of material point. A material point located on a surface has two
degrees of freedom.
A material point compelled to remain on a curve has only one degree of freedom.
He can only move along the curve.
Finally, a fixed material point has no degree of freedom.

5.2. Equilibrium of the free material point

The necessary and sufficient condition for a free material point, which is at rest or
in rectilinear and uniform motion, to remain in the same condition under the action of a
→ →
mechanical system of n competing forces F i , i = 1, n , is that the resultant of the forces R to
be zero. This condition results from the application of inertia and force action principles.
Consequently, the condition of equilibrium of the free material point is:
→ n → →
R= ∑F
i =1
i =0 (5.1)

Projecting this vectorial relation onto the axes of a Cartesian system Oxyz, the
scalar equations of equilibrium are obtained:
n n n
R x= ∑
i =1
Fi x = 0 , R y = ∑ i =1
Fi y = 0 , R z = ∑F
i =1
iz =0 (5.2)

If the system of forces is plane and if we denote by Oxy the plane of forces, the
scalar conditions of equilibrium are of the format:
n n
R x= ∑F
i =1
ix = 0 , R y= ∑F
i =1
iy =0 (5.3)

62
Theoretical mechanics – theory and applications

5.3. Equilibrium of the constrained material point. Principle of connections.


Classification of connections.

5.3.1. Principle of connections

Let a material point M be in equilibrium on a surface (S) and acted upon by a



system of external forces whose resultant is R (Figure T 5.1). Unlike the free material
→ →
point case, when, R = 0 was required to ensure the equilibrium, such condition must not
be observed for the material point constrained due to existence of the connections that
exert on the material point certain mechanical constraints represented by the forces of
connection (reactions). To solve the problem of the connected material point the principle
of connections is used, whose statement in the general case is:
Principle of connections: Any connection may be suppressed and substituted by
appropriate mechanics elements (forces, moments).

In the case of a constrained material point, the connection is substituted by a



reaction R ’, so that the condition necessary and sufficient for a material point subject to
connections to be in equilibrium is that the resultant of the forces directly applied and of
the connection force is null:
→ → →
R + R '= 0 (5.4)
→ →
i.e. the reaction (connection force) R ’ is directly opposite to the resultant R of the directly
applied forces.
(S)

M 
• R′

R

Figure T 5.1

Note i) When the material point is subject to more connections, then R ’ is the resultant of
the forces of connection corresponding to each single connection.

5.3.2. Classification of the material point connections

The material point connections are three in number, namely: supporting on a


surface, supporting on a curve (plane or space) and attaching to wires.
They can be:
a) with friction – if support surface or curve belong to real bodies, which have a
rough surface resisting the motion and giving rise to friction forces.
b) frictionless – when it is assumed that the bearing surface or curve belong to
the ideal body, perfectly smooth, which do not give rise to friction forces.

63
Theoretical mechanics – theory and applications

5.4. Equilibrium of the material point subject to frictionless connections

5.4.1. Equilibrium on a frictionless (smooth) surface

Let a material point M be supported on a surface (S) and acted upon by directly
→ → →
applied forces whose resultant is R and by the reaction R ’ (directly opposite to R ).

Resultant R (Figure T 5.2) resolves in:

- the normal component R n (about normal in M at (S));

- the tangential component R t (in the tangent plane (P) at (S)).

normala

 (S)
(P)   N
N R′

 
Rt M


T • M
Rn

R

(S)
R

Figure T 5.2 Figure T 5.3



In its turn, the reaction R ’ resolves about the same directions into the components
→ → →
N and T . Force R n is trying to remove the point M from the surface (S). Its effect is

cancelled by the force N , called normal reaction:
→ → →
R n+ N = 0 (5.5)

Force R t determines motion of the point M on the surface (S). It is resisted by the

component T , called friction force:
→ → →
R t+ T = 0 (5.6)

Since the surface is considered smooth (no friction) the force T is zero so that
using equation (5.6) we may write that:
→ → → →
R = R n+ R t= R n (5.7)
i.e. in order to achieve the equilibrium on the surface (S), the resultant of the directly
applied forces must be conducted about the normal to the surface at that point. (Figure T
5.3).

64
Theoretical mechanics – theory and applications

The vectorial equation of equilibrium is:


→ → →
R+ N = 0 (5.8)
and its projections on the axes of a Cartesian reference are as follows:

R x+ N x= 0 , R y+ N y= 0 , R z+ N z= 0 (5.9)

Should surface (S) is given by the equation f(x, y, z) = 0, since the directing
∂f ∂f ∂f
parameters of the normal are , , , then the normal reaction shall have the
∂x ∂y ∂z
expression:
→  ∂ f → ∂ f → ∂ f →
N = λ ⋅ grad f = λ ⋅  i+ j+ k (5.10)
∂x ∂y ∂ z 
and the vectorial equation of equilibrium is:
→ →
R + λ ⋅ grad f = 0 (5.11)

The scalar equations of equilibrium


∂f ∂f ∂f
R x+ λ = 0 , R y+ λ = 0 , R z+ λ =0 (5.12)
∂x ∂y ∂z
form along with the equation of the surface f(x, y, z) = 0 a system of four equations with
the unknown quantities λ , x, y, z .

5.4.2. Equilibrium on a frictionless (smooth) curve

Let a material point M be supported on a curve (C) acted upon by directly applied
→ → →
forces whose resultant is R and by the reaction R ’ (Figure T 5.4). Resultant R resolves
in:

- component R t (about the tangent M to curve ( C ));

- component R n (in the plane normal at M to curve ( C )).
→ → →
Reaction R ’ resolves about the same directions into the components T and N .
Relations (5.5) and (5.9) as well as the comments associated thereto remain valid in this
case, too. (Figure T 5.5).
If the curve is given as an intersection of two surfaces:

(S1 ): f1 (x, y, z ) = 0 , (S 2 ) : f 2 (x, y, z ) = 0 (5.13)


then the plane normal to the curve ( C ) can be obtained with the help of the normals to the
two surfaces, the normal reaction being of the format:
→ → →
N = N 1 + N 2 = λ1 ⋅ grad f1 + λ 2 ⋅ grad f 2 (5.14)

65
Theoretical mechanics – theory and applications

From (5.8) and (5.13) the following scalar equations of equilibrium are obtained:
∂f ∂f ∂f ∂f
R x + λ 1 1 + λ2 2 = 0 , R y + λ 1 1 + λ2 2 = 0
∂x ∂x ∂y ∂y
∂f ∂f
R z + λ 1 1 + λ2 2 = 0 (5.15)
∂z ∂z
which along with the equations (5.13) of the surfaces compose a system of five equations
with the unknown quantities λ 1, λ 2 , x, y, z .

Plan
normal Plan
normal


R′  
N N
 
T M Rt tangenta
tangenta M

 ( C) 
Rn  Rn ( C)
R

Figure T 5.4 Figure T 5.5

5.5. Equilibrium of the material point subject to connections with friction


Laws of dry friction. Friction cones.

5.5.1. Laws of dry friction (Coulomb’s Laws)

Let a material point M be on a surface or a curve (rough). In order to highlight the


friction force Coulomb used the following experience performed by a machine called
tribometer:

A body assimilated to a material point of weight G is placed on a horizontal plane

and acted upon by a force F (by means of weights placed on a tray) which can vary

continuously from 0 to ∞ (Figure T 5.6). It is found that up to a certain value F max the
body does not move.

That is a proof that reaction R ' is inclined by an angle α with the normal since,

otherwise, under the action of the force F the body would move, no matter how small the
→ →
intensity of this force is. Reaction R ' resolves into its components: normal reaction N and

sliding force of friction T .

66
Theoretical mechanics – theory and applications

 
R′  R′max 
N
N
α 
F ϕ 
 Fmax
T 
Tmax

G 
G

G 
G

(a) (b) (c)


Figure T 5.6

Force T acts, for the general case of a certain bearing surfaces, in the plane
tangent to the surface and opposes the trend of movement.
Its magnitude is:
→ →
T = N ⋅ tg α (5.16)

→ →
Figure T 5.6.c shows the limit case when forces F and T take the maximum

values in equilibrium (any greater values for F leading to breakage of the equilibrium)
Angle α also has a maximum value which we will denote by ϕ and shall call angle of
friction. We can write:
→ →
T max = N ⋅ tg ϕ (5.17)

And since α ≤ ϕ , we find that:


→ →
T ≤ N ⋅ tg ϕ (5.18)

Coulomb stated the following laws, called laws of dry friction (see also equation
(5.17):
1) The size of the maximum sliding force of friction is proportional to the size of
normal reaction;
2) The size of sliding force of friction depends on the nature and condition of surface
of bodies in contact;
3) The size of sliding force of friction is dependent neither of the relative velocity of
movement of two bodies in contact nor the contact surface size.
Based on the above laws the sliding force of friction is expressed as:

→ → →
T ≤ T max = µ ⋅ N (5.19)

67
Theoretical mechanics – theory and applications

where µ = tg α is called sliding coefficient of friction and is a dimensionless quantity that


depends on the nature and the condition of the surfaces in contact.
Coulomb considered that the friction forces arise from the existence of asperities at
the surface of bodies that interlock when two bodies come in contact. When one of the
body moves (or both of them) asperities are sheared off, sliding friction force being itself
the force that opposes shearing.
Notes: After the experiments of Coulomb, by extension thereof, some corrections to these
laws have been brought. Thus, it is found out that the coefficient of friction decreases as
the velocity of the bodies in contact increases, its variation being shown in Figure 5.7. The
value µ 0 of the coefficient of friction at zero velocity is called coefficient of adhesion.

Also, for larger values of the reaction R ' the size of the force of friction does no longer
vary linearly with the size of the reaction.

µc

O v
Figure T 5.7

5.5.2. Friction cones

Let us further analyze the geometrical interpretation of the equilibrium of the


material point with friction. Considering the point supported on a surface and changing the
→ → →
direction of the force F in the tangent plane, the reaction R ' and the resultant R ,
respectively will describe a cone, called cone of friction, which has the vertex at point M
and the angle at vertex 2ϕ (Figure T 5.8). The material point is in equilibrium when the

reaction R ' finds within the cone of friction or, in the limit, on its surface.


R′ ϕ 
R′ tangenta
(S)
M
π
−ϕ
2 M

R

( Γ) R

Figure T 5.8 Figure T 5.9

68
Theoretical mechanics – theory and applications

In the case of a material point supported with friction on a curve ( Γ ), because


through that curve an infinite number of surfaces are passing and each one has a cone of
friction, the extreme generatrices thereof will describe complementary cones of friction
(Figure 5.9). Such cones have the tangent to the curve at the considered point as the
π 
symmetry axis and the angle at the vertex 2  − ϕ  . The material point is in equilibrium
2 

when the reaction R ' is outside the complementary cones of friction or, in the limit, on the
surface thereof.

5.6. Solved problems

R 5.1) Ball M , of weight G , is in equilibrium on a smooth plane inclined at the angle α


with the horizontal, attached to two wires parallel to the inclined plane, making the
angles β and γ with the largest slope line of the plane. (Figure R 5.1.1). Determine the
tension in the wires and the reaction of the plane.
y y z

x 
A1 A2 A1
γ
A2 ⊗ N
β γ β
•M
 
α T1 T2
x α α

M M ⋅ z G

(a) (b)

Figure R 5.1.1 Figure R 5.1.2.

→ →
Solution: Denote with T 1 and T 2 tensions in the wires MA 1 and MA 2 , respectively, and

with N the normal reaction (see Figure R 5.1.2). For equilibrium the condition
→ → → → →
G + T 1+ T 2 + N = 0 is imposed. Projecting this vectorial equation onto the axes of the
Cartesian reference Mxyz, the resulting scalar equations are:

∑ X i = T 2 sin γ − T 1 sin β = 0
∑ Y i = −G sin α + T 1 cos β + T 2 cos γ =0
∑ Z i = N − G cos α = 0
Solving this system we obtain:
sin α sin γ sin α sin β
N = G cos α , T 1= G , T 2= G
sin( β + γ ) sin( β + γ )
.

69
Theoretical mechanics – theory and applications

R 5.2. Let the smooth curve of equation y = x be related to the system of axes Oxy, the
plane of the curve rendering the angle α with the horizontal plane. A ring of weight G
may glide onto this curve being acted upon by the force F parallel to Ox, as well. (Figure
R 5.2.1). Determine the position of equilibrium of the ring on the curve and the reactions
N1 (in the plane of the curve) and N 2 , respectively, (perpendicular on the plane of the
curve).

y= x normala tangenta z y
N1
β N 2 N1 cos β
M β
y  F
F •
 G sin α
G α α α

z• O x 
O x0 x
G

(a) (b)

Figure R 5.2.1 Figure R 5.2.2.

Solution: Let x0 be the abscissa of the point of equilibrium and


1
β = arctg f ' ( x 0 ) = arctg the angle rendered by the tangent to the curve y = x
2 x0
→ → → → →
with axis Ox at this point (see Figure R 5.2.2). For equilibrium: G + N 1+ N 2 + F = 0 . On
the axes of the reference Oxyz we have the following scalar equations of equilibrium:
 G2
x 0= sin 2 α
 ∑ X i = F − N 1sin β = 0
2
 4F
 
 ∑ Y i = N 1 cos β − G sin α = 0 ⇒  N 1= F 2 + G 2 sin 2 α .
 
 ∑ Z i = N 2 − G cos α = 0  N 2 = G cos α


R 5.3. A ring of weight G, supported with friction (coefficient of friction µ ) onto a vertical

circle is acted upon with the horizontal force F , of magnitude G. Determine the positions
of equilibrium of the ring, given by the angle θ (Figure R 5.3).
Solution: a) The trend of motion of the ring towards point A:
∑ X i = T + F cos θ − G sin θ = 0 (1)
∑ Y i = N − F sin θ − G cos θ = 0 (2)
For equilibrium: T ≤ µ N (3)

(1) ⇒ T = G sin θ − F cos θ , ( 2) ⇒ N = F sin θ + G cos θ

70
Theoretical mechanics – theory and applications

F+µG
(3) ⇒ tg θ ≤ (4)
G− µ F

b) The trend of motion of the ring towards point B:


Reversing the sense of the force of friction we obtain:
F−µG
tg θ ≥ (5)
G+ µ F
1 − µ 1 + µ 
Since G = F, from (4) and (5) we find that tg θ ∈ 
 1 + µ 1 − µ 
, .

O

M F
θ

G

Figure R 5.3

R5.4. Determine the sliding coefficient of friction for which a material point is in
 R
equilibrium on the surface of equation x 2 + y 2 = R ( R − z ) , at position M  x, y,  , under
 2
→ → 2G ____  R
the action of the weight G and of the elastic force a F = ⋅ MA , where M  0, 0,  .
R  2
Solution: The equilibrium on a rough surface is achieved if the line of action of the
resultant of the forces applied is within the cone of friction. For a point located on the
surface f(x, y, z) = 0 the condition to be met is:
∂f ∂f ∂f
X⋅ +Y ⋅ +Z⋅
∂x ∂y ∂z 1
≥ (1)
 ∂f  2
 ∂f  2
 ∂f  2
1 + µ 2
X 2 +Y 2 + Z2 ⋅   +  + 
 ∂x   ∂ y   ∂ z 

where X, Y, Z are the components of the resultant R . But:
∂f ∂f ∂f
f ( x, y , z ) = x 2 + y 2 − R ( R − z ) = 0 ⇒ = 2x , = 2y , =R (2)
∂x ∂y ∂z
→ → → → 2G  → → 2G 2G
R = G + F = −G k + − x i − y j  ⇒ X = − x ,Y = − y , Z = −G (3)
R   R R
1
From (1), (2), (3) and the equation of the surface we get µ ≥ .
2 2

71
Theoretical mechanics – theory and applications

5.7. Problems proposed

5.7.1. Conventional tests

TC 5.1) The load Q = 100 daN is supported by the beam AO, hinged at O and inclined at
an angle of 45 0 with an upright wall and with two chains BA and CA equal in length, laid
 ∧   ∧ 
horizontally ( mCBA = m ACB  = 45 0 ). Determine the efforts in the beam AO and in the
   
two chains (Figure TC 5.1).
z

x 2 + y2 − z = 0
B
x+y+z =0
D 450
A(−1,−1,2) Γ
C 450
A
450
O Q O y

Figure TC 5.1 Figure TC 5.2

TC 5.2) Let us consider the smooth curve ( Γ) given as an intersection between the
paraboloid of revolution of equation f ( x, y, z ) = x 2 + y 2 − z = 0 and the plane (P) of
equation g ( x, y, z ) = x + y + z = 0 . Determine the positions of equilibrium of a point on this
→ →
curve if the former is subject to the action of its own weight G = − mg k and of the force
→ → → →
F = mg i + j − k  (Figure TC 5.2).
 
TC 5.3) On the faces ABB’A’ and ACC’A’ of the prism ABCA’B’C’
 ∧   ∧ 
( m BAC  = m ACB  = 45 0 ) there are the bodies of weight P and Q, respectively. The
   
bodies are connected through a thread passing over the frictionless pulley block at B (the
thread is directed over the parallels to AB and BC). If Q = 2P and the coefficient of friction
between the bodies and the prism is µ , determine the minimum and maximum values of
the inclination angle α of the prism in order to achieve the equilibrium. (Figure TC 5.3).

TC 5.4) Determine the positions where a heavy material point can remain at rest on a
rough sphere of equation x 2 + y 2 + z 2 − r 2 = 0 , the coefficient of friction being µ .

72
Theoretical mechanics – theory and applications

C

⋅ Q
⋅ A

 B C
P 450 600
F
300 B

100N
α

A 100 N

Figure TC 5.3 Figure TG 5.1

5.7.2. Multiple-choice test

TG 5.1) Determine the tension in the cables AB and BC loaded as in Figure TG 5.1.
Pulleys in E and F are frictionless.
( ) ( )
a) T AB = 100 2 2 − 6 , TBC = 100 2 − 3 ; b) T AB = 100(1 + 3 ), TBC = 200( 6 − 2 ) ;
c) T AB = 100(2 2+ 6 ), TBC = 100(2 + 3 ) ;
d) d) T AB = 100( 2 + 3 ), TBC = 100( 6 + 2 ) .

TG 5.2) A sphere of weight P lays on a smooth cylindrical surface of radius r, being


hanged on a thread at the fixed point A. Knowing the length l of the thread and the angles
α and β , determine the tension in the thread and the reaction of the cylindrical surface
(Figure TG 5.2).
r r sin α sin β
a) T = P sin α , N = P sin β ; b) T = P ,N = P
sin(α + β ) sin(α + β )
;
l l
sin(α − β ) sin α l r
c) T = P ,N = P ; d) T = P sin β , N = P sin α .
sin(α + β ) sin β r l

A

β
l

α •
r

Figure TG 5.2

73
Theoretical mechanics – theory and applications

5.8. Indications and answers

TC 5.1) Denote by S 1 and S 2 the tensions in the chains and by S 3 the effort in the beam
OA. The equations of equilibrium on three perpendicular directions are:
S 1 cos 45 0 + S 2 cos 45 0 − S 3 sin 45 0 = 0
S 1sin 45 0 − S 2 sin 45 0 = 0
−Q+S 3 cos 45 0 = 0
S3
and the solution is S1 = S 2 = = 70,5 daN .
2
→ → →
TC 5.2) The vectorial equation of equilibrium G + F + λ 1⋅ grad f + λ 2 ⋅ grad g = 0 is


=N
projected on the axes of the Cartesian reference Oxyz:
∂f ∂g
∑ X i = mg + λ 1⋅ ∂ x + λ 2⋅ ∂ x = 0
∂f ∂g
∑ Y i = mg + λ 1⋅ ∂ y + λ 2⋅ ∂ y = 0
∂f ∂g
∑ Z i = − mg − mg + λ 1⋅ ∂ z + λ 2⋅ ∂ z = 0 .
A system of five equations is obtained, with the unknown quantities λ 1, λ 2 , x, y, z :
f ( x, y , z ) = x 2 + y 2 − z = 0 , g ( x, y , z ) = x + y + z = 0 ,
m g + 2 x λ 1+ λ 2 = 0 , m g + 2 y λ 1+ λ 2 = 0 , − 2 m g − λ 1+ λ 2 = 0 ,
with the solution:
a) x 1= y 1= z 1= 0 , λ 1= −3 m g , λ 2 = − m g ;
b) x 2 = y 2 = −1, z 2 = 2 , , λ 1= 3 m g , λ 2 = 5 m g .

TC 5.3) The trend of motion of the system of the two bodies towards point C is studied
(Figure TC 5.3.2).
x y
 
T  
y  T ⋅ T2 N2
N2
π
 −α
4
 P Q
T1 x
π
−α
4
α

Figure TC 5.3.2

74
Theoretical mechanics – theory and applications

The scalar equations of equilibrium for the body of weight P are:


π 
∑ X i = T − T 1− P cos 4 − α  = 0 (1)

π 
∑ Y i = N 1− P sin 4 − α  = 0 (2)

T 1≤ µ N 1 (3)
and for the body of weight Q:
π 
∑ X i = −T − T 2 + Q sin
4
−α= 0

(4)

π 
∑ Y i = N 2 − Q cos 4 − α  = 0 (5)

T 2≤ µ N 2 (6)

1− 3 µ not
From this system we find that α ≥ α = arctg
. Studying also the trend
3+ µ
min

of motion of the system of two bodies towards A we deduce the maximum value of the
not 1+ 3 µ π  1 − tg α
angle α : α ≤ α max = arctg . The fact that G = 2 P and tg − α  = has
3− µ 4  1 + tg α
been used. The equilibrium is attained if α ∈ [α min , α max ] .

R ⋅ grad f
1
TC 5.4) Impose the condition →
≥ ⇔
R ⋅ grad f 1+ µ 2

∂f ∂f ∂f
R x⋅ + R y⋅ + R z⋅
∂x ∂y ∂z 1
≥ ,
 ∂ f 2  ∂ f 2  ∂ f 2 1+ µ 2
R x+ R y+ R z ⋅ 
2 2 2  +  + 
 ∂ x   ∂y   ∂ z 
where R x = R y = 0, R z = G , f ( x, y, z ) = x 2 + y 2 + z 2 − r 2 . It is obtained that:
 
r
z ∈ , r  , x, y = arbitrary.
 1 + µ 2 

TG 5.1) Tensions in the threads BE and BF are each equal to 100. The equations of
equilibrium in horizontal and vertical directions are:
T AB cos 45 0 − T BC cos 60 0 + 100 cos 30 0 − 100 = 0
T AB sin 45 0 − +T BC sin 60 0 − 100 sin 30 0 = 0 .
Correct answer: a).

75
Theoretical mechanics – theory and applications

→ → → →
TG 5.2) The vectorial equation of equilibrium, P + N + T = 0 , is projected onto the
vertical and onto the horizontal and obtain the scalar equations:

T cos β + N cos α − P = 0 , T sin β − N sin α = 0 .


Correct answer: b).

76
Theoretical mechanics – theory and applications

6. Equilibrium of the rigid


6.1. Equilibrium of the free rigid

The free rigid is one that can occupy any position in space, its position is only
determined by the forces acting upon it.
At the reduction of the system about a point O it was shown that if
→ → →
R = M O = 0 the given system of forces is equivalent to zero and according to the
principle of inertia, it does not change the mechanical condition of the body on which it
acts. Thus, if it was in equilibrium, it will continue to remain at rest upon application of
such a system of forces.
It follows that for a rigid to remain in equilibrium under the action of a system of
forces, it is necessary and sufficient that at every point O thereof the reduction torsor has
both components equal to zero:
→ → → →
R= 0 , M0 = 0 (6.1)
Projecting the two vector equations on the axes of a Cartesian system Oxyz, six
scalar equations of equilibrium are obtained:
n n
Rx = ∑ Fi x = 0 , M Ox = ∑ Mix = 0
i =1 i =1
n n
Ry = ∑ Fi y = 0 , M Oy = ∑Miy = 0 (6.2)
i =1 i =1
n n
Rz = ∑ Fi z = 0 , M Oz = ∑ Miz = 0
i =1 i =1
In the case of some particular systems of forces the number of scalar equations of
equilibrium decreases. Thus, for a system of coplanar forces acting in the plane Oxy, the
number of scalar equations of equilibrium decreases to three:
n n n
Rx = ∑ Fi x = 0 , R y = ∑ Fi y = 0 , M Oz = ∑ Miz = 0 (6.3)
i =1 i =1 i =1
The equilibrium of the rigid is achieved for a given configuration thereof relative
to a reference, i.e. for well-defined positions of points of the rigid relative to the reference.
Thus, to know a rigid position in relation to a Cartesian reference system, it is enough to
know the position of three of its points, and so to know nine Cartesian coordinates. Since
only six of the coordinates are independent (there are three relations between coordinates,
which express that the distance between points is constant) the free rigid solid has six
degrees of freedom in space (motions on Ox, Oy, Oz and spins around the same axes).

77
Theoretical mechanics – theory and applications

6.2. Equilibrium of the rigid solid subject to constraints. Generals

Constraints to which a rigid may be subject are more complex than those a
material point is subjected to, but no matter how complex they are, the equilibrium is
studied the same way. The constraints of the body are removed one by one and replaced in
accordance with the principle of connections, with connection forces or couples. Under the
action of the given forces and of the connection forces the rigid may be considered free.
Reducing the forces given by the connection about a point O, arbitrarily chosen,
→ →  → →' 
   
and denoting with R , M O the torsor of the given forces and with  R ' , M O  the torsor
   
   
of the reaction, the vector equations of equilibrium are in the form:
→ → → → → →
R + R'= 0 , M 0 + M O' = 0 (6.4)

Projecting these equations on the axes of a Cartesian reference Oxyz, we obtain six
scalar equations of equilibrium:
R x + Rx ' = 0 , R y + R y ' = 0 , R z + Rz ' = 0
(6.5)
M Ox + M Ox ' = 0 , M Oy + M Oy ' = 0 , M Oz + M Oz ' = 0 .

A problem of equilibrium of a rigid solid subjected to constraints is determined, if


the total number of scalars unknown is six, and if it is greater, then the problem is
indeterminate, existing the possibly of an infinity of equilibrium positions or an infinity of
values for the connection forces or both.
Any motion of a rigid solid can be decomposed in general into two basic motions:
- a translation motion;
- a rotary motion around an axis.
If the constraint prevents translation, then the reaction will be a force in the
direction of the prevented motion but opposite in sense. If the constraint prevents
revolution about an axis, then the constraint reaction will be a couple of forces located in a
plane perpendicular to that axis, opposite to the rotation prevented. We will use these
observations in what it comes further.

6.3. Equilibrium of the rigid subject to frictionless constraints

The constraints of the rigid are the support (simple support), the joint, the
embedding and attachment to wires. In the study of the rigid constraints, two aspects are
considered, namely:
- The geometric aspect, concerning the number of degrees of freedom left to a rigid
upon application of the constraint;
- The mechanical aspect, concerning the mechanical elements (forces, moments)
replacing the constraint. We will study below the ideal constraints, that is friction
will be neglected.

78
Theoretical mechanics – theory and applications

6.3.1. Simple support

The simple support represents the constraint whereby a point of the rigid O is
forced to remain continuously on a given surface or curve.

If O(xO , yO , zO ) and the equation of the support surface is f(x, y, z) = 0, then the
condition that the point O belongs to surface f (xO , yO , zO ) = 0 shall be another relation
between the nine coordinates of the three non-collinear points fixing the rigid in space.
Consequently, a simple support suppresses a rigid a degree of freedom so that a rigid with
a simple support has five degrees of freedom.
Normala

(S)

N
Plan tangent


 Mt
Rt O 
M0

Mn
  f ( x , y, z) = 0
R Rn

Figure T 6.1
To study the forces that appear at O (Figure T 6.1) the elements of the torsor of the
→ → 
external forces  R , M O  are broken down in pairs of two:
 
 
→ → → → → →
R = Rn+ Rt , MO = M n+ Mt (6.6)
→ → →
Force R t and moments M n , M t tend to put the rigid (S) into motion. The lack
of friction makes it impossible to stop these motions, so that, for the solid to remain in
equilibrium it is necessary that:
→ → → →
Rt =M n =Mt = 0 (6.7)
It results that:
→ → → → →
R = Rn ≠ 0 , MO = 0 (6.8)
→ →
According to the principle of action and reaction, a constrain force R ' = N , equal

and directly opposite to R n occurs at O, so that we can state that a frictionless support is
→ →
replaced by a force R ' = N (normal reaction), directed by the normal common at the
point of contact.

79
Theoretical mechanics – theory and applications

6.3.2. Joint

Joint is a rigid solid constraint whereby a point O thereof is forced to remain


permanently in a fixed point.

Joint can be plane (cylindrical) when the rigid is acted upon by a system of plane
forces or spatial (spherical) if the solid is acted upon by a system of spatial forces.

a) Case of the spherical joint


Forcing the point O( xO , yO , zO ) to stay fixed, three geometrical conditions are
imposed (xO , yO , zO = cons tan t ) which decrease the number of degrees of freedom of the
rigid from six to three, such only three revolutions being allowed about three axes
concurrent at O. To study the restraint forces, the reduction torsor  R, M O  of the external
→ →

 

forces at O are considered. Moment M O tends to rotate the body and since there is no
→ → → ' →
friction to resist such tendency it is required that M O = 0 (hence M O = 0 ). It follows that

the joint reaction is a force R ' of a certain magnitude and direction. (Figure T6.2).

Resolving the force R ' about the axes of the Cartesian reference Oxyz into the
components R 'x , R 'y , R 'z it results that a spherical joint introduces three scalar unknowns
(Figure T6.3). This number equals the number of lost degrees.
z

(S) (S)

 R′z
R′

O
R′y
 O
 y
R
R′x

Figure T 6.2 Figure T 6.3

b) Case of plane hinge


Forcing point O( xO , y O ) to remain fixed, two geometrical conditions are imposed
(xO , yO = cons tan t ) which decrease the number of degrees of freedom from three to one,
the point being allowed a single revolution around an axis perpendicular to the plane of the

forces and that passes through O. In this case, too, the joint is replaced with a reaction R '
but this one is located in the plane of forces and resolves about the directions Ox and Oy

80
Theoretical mechanics – theory and applications

into the components R 'x and R 'y (Figure T 6.4). It follows that a plane joint introduces two
scalar unknowns (number equal to the number of lost degrees)

y M ′0
(S)
(S)
R′y 
 R′
R

O M0
R′x x

Figure T 6.4 Figure T 6.5

6.3.3. Embedment

The embedment is the restraint by which a body is rigidly fixed (stuck) into
another body so that no motion is allowed.

The body has no degree of freedom left. From the mechanical point of view, the
→ → ' 
embedment is equivalent to a torsor  R ' , M O  , where O is the center of weight of the cross
 
area at the constraint. (Figure T 6.5). The two components of the torsor are arbitrary
→ → → ' → 
vectors  R ' = − R, M O = − M O  and each of them may resolve in three components onto
 
the axes of a Cartesian reference Oxyz (Figure T 6.6). An embedment introduces six scalar
unknowns: R 'x , R 'y , R 'z , M Ox
' '
, M Oy '
, M Oz (number equal to the number of lost degrees).
In the case of an embedded body acted upon by plane forces, the number of scalar
unknowns is three: R 'x , R 'y , M Oz
'
(Figure T 6.7).

z y
M′Oz
(S)
S
R′z R′y
A
R′y
O
R′x y
M′Oz O
R′x x
M′Oz
M′Oz
x

Figure T 6.6 Figure T 6.7 Figure T 6.8

81
Theoretical mechanics – theory and applications

6.3.4. Attachment to wires

The attachment to wires is treated in the case of the rigid, same as for the material
point, i.e. such connection is replaced by a force that is introduced along the cut wire so

that it strains the portion of wire left tied up to the rigid (Figure T 6.8). Force S is called
tension in string.

6.4. Equilibrium of rigid subject to friction constraints

6.4.1. Types of friction. Equations of equilibrium.

Subchapter 6.3 treated the equilibrium of the rigid with ideal constraints. In reality,
the connections of the bodies are always accompanied by friction. Physical explanation is
that in reality the bodies are deformable and enter into contact not in a single point O but

an entire connection surface on which the connection forces p i are very difficult to
establish (Figure T6.9). Contact surfaces have protuberances which, under the action of
forces acting on bodies, interlock and undergo deformation.

 (S) 
F1 F2


Fn
O
• Ai



pi

Figure T 6.9

Let a solid rigid (S) be acted upon by the external forces F i , i = 1, n . The torsor of
the external forces at the theoretical contact point O (should bodies would not deform) is
→ →
composed of the resultant force R and resultant moment M O . The torsor at O of the

contact forces p i applied at A i is:
→ →
' ∑
 R ' = pi
τO  '
→ → →

M O = OA i × p i

The equations of equilibrium are:


→ → → → → →
R + R '= 0 ' = 0
, MO+ M O (6.9)

82
Theoretical mechanics – theory and applications

Each of the components of the reduction torsor at O of the external forces


decomposes about the normal at O to the contact surface and a line in the plane tangent at
O to the support surface (S 2 ) :
→ → → → → →
R = Rn+ Rt , M 'O = M p + M t (6.10)
' of the
Identically, (Figure T 6.10) is proceeded with the elements of the torsor τ O
connection forces:
→ → → → → →
R '= N + T , M 'O = M p + M r (6.11)
The equations of equilibrium, projected on the normal direction and on the tangent
plane are written as:
→ → → → → → → → → → → →
Rn+ N = 0 , Rt+ T = 0 , M n+ M p = 0 , Mt+Mr = 0 (6.12)

(S1 )

normală în O

 
N R′
plan tangent in O

M′O

Mp

Mr

T


 Mt
Mn

MO
 
R Rn

(S2 )

Figure T 6.10

Let us analyze the role played by the individual component:



• Force R n tends to move the body in the direction of the normal On to the

support surface. This motion is prevented by the normal reaction N ;

• Force R t tends to move the body in the plane tangent at O to the support
surface. This movement is called sliding and it is prevented by the sliding friction

force T ;

83
Theoretical mechanics – theory and applications


• The couple of moment M n tends to rotate the body around the normal

On. This motion is called pivoting and it is prevented by the couple of moment M p
called pivoting friction couple;

• The couple of moment M t tends to rotate the body around an axis in the
plane tangent to the contact surface. This motion is called rolling and it is resisted by

the couple of moment M r , called rolling friction couple.

6.4.2. Sliding friction

Let us consider a rigid acted upon by a system of external forces whose reduction
torque about the theoretical point of contact with the support surface is only composed of
→ → →
force R = R n + R t (Figure T 6.11). According to the action and reaction principle, the
→ → → →
force R is resisted by the reaction R ' = N + T . This force is inclined at the angle α with
the normal direction, where α is given by the relation:
→ →
T = N ⋅ tg α (6.13)

Normala

 
R′ R′max
 
 R′ N
N ϕ
α
 O
Rt T O  
T Tmax
Plan
  tangent
R Rn

Figure T 6.11

In the case of equilibrium in the limit, the angle α reaches the maximum
value α max = ϕ :
→ →
T max = N ⋅ tg ϕ (6.14)

not
Denoting µ = tg ϕ , where µ is the sliding friction coefficient, we get:

84
Theoretical mechanics – theory and applications

→ → →
T ≤ T max = µ ⋅ N (6.15)

The laws of dry friction (Coulomb) presented in the case of the equilibrium of the
material point with friction as well as the comments on the coefficient of sliding friction µ
will also apply in this case.

6.4.3. Rolling friction

Let us consider a rigid in equilibrium acted upon by a system of forces whose


→ → → → →
torsor at O is composed of R = R n + R t and M O = M t (Figure T 6.12). According to
the action and reaction principle, the torsor of the connection forces at the same point is
→ → → → → →
composed of R ' = N + T and M 'O = M r . The moment M t tends to produce rolling of

the body while it is resisted by the rolling friction moment M r . This situation is met in
practice in the case of car wheels, and of bearing balls or rolls.

 (S)
R′

   
Mo = M t M′o = M r
O

plan
tangent

R normala

Figure T 6.12

Thus, in Figure T 6.13 (a) let us consider a wheel acted upon by the external forces
→ → → →
G and F . In figure T 6. 13 (b), forces G and F have been replaced by the reduction
torsor about the theoretical point of contact O. The wheel and the rolling way are
deformable and hence, the contact shall take place on a surface rather than a point. At the
individual point of contact a reaction appears, which may be decomposed on the normal
→ →
and tangential directions, as n i and t i . These components are replaced with their
→ →
resultants N and T .
Situation in Figure T 6.13 (b) is determined by the fact that the contact area is
asymmetrically related to the center plane, larger in the direction the wheel tends to move
towards. For the equilibrium in the limit, the situation in Figure T 6.13 (c) is obtained.

85
Theoretical mechanics – theory and applications

   
C F R C R C C
• • • R

  b
ti G 
•   •    M0
a M0 M0 T T M
  
Δ r
N
N N

(a) (b) (c) (d)


Figure T 6.13

Distance s, that represents the maximum distance a, whereby the normal reaction

N about the vertical direction of the point O is displaced, is called coefficient of rolling
→ →
friction. Distance b is negligible. Performing reduction of the force system N and T about
→ →
point O, the situation in Figure 6.13 (d) is obtained. In conclusion, forces N and T as

well as the rolling friction moment M r resisting the rolling trend, of magnitude

→ → →
M r ≤ M r , max = s ⋅ N (6.16)

occur as reactions.
The coefficient s is expressed in meters and its value (determined experimentally)
depends on radius of the wheel and on the nature of the materials the wheel are made from.

6.4.4. Pivoting friction

Let us consider a rigid in equilibrium on a surface and acted upon by a system of


→ → → →
external forces whose torsor about point O ∈ (S ) is R = R n and M O = M n . The torsor
→ → → →
of the connection forces about the same point is composed of R ' = N and M 'O = M p
(Figure T 6.14). Equilibrium requires that:
→ → → → → →
Rn+ N = 0 , M n+ M p = 0 (6.17)

Experience shows that M p , called pivoting friction moment, may not exceed a
certain limit value, hence, the maximum value of the pivoting friction is given by:
→ →
M p, max = µ k N (6.18)

86
Theoretical mechanics – theory and applications

where µ is the sliding friction coefficient (dimensionless) and k is a factor that depends on
the shape of the surfaces in contact (has the dimensions of a length).

(S1 ) normala


N

Mp

(S2 ) O

Mn

Rn

Figure T 6.14

In conclusion, the pivoting tendency is resisted by a pivoting friction couple


ranging from zero to a maximum value, equal to the product between a sliding friction
coefficient µ (established empirically), a one-dimension factor k depending on the shape
and dimensions of the surfaces in contact, and the magnitude of the normal reaction:

→ → →
M p ≤ M p ,max = µ k N (6.19)

6.4.5. Friction in joints and bearings


Besides the friction cases already presented, that referred to the simple support, in
engineering other important cases involving friction are found. Two such situations shall
be presented below. Wheels of a car are attached to shafts or axles, which in turn rest on
bearings (Figure T 6.15). The portions of shafts or axles entering in contact with the
bearings are called journals. During the rotation of the shaft, there is a sliding or a rolling
accompanied by friction, between the journal and the bearing.

Tendinţa de mişcare
Fus
Fus Rulment

lagăr


Mf
r

• O

R′

lagăr lagăr Fus


A

Figure T 6.15 Figure T 6.16

87
Theoretical mechanics – theory and applications

Friction in a bearing acts as a friction couple. Experience shows that this moment,

denoted M f and called bearing (joint) friction moment may not exceed a certain limit,
hence the maximum value of the bearing friction moment is given by:
→ →
M f ,max = µ 'r R ' (6.20)

where µ ' is the (dimensionless) friction coefficient in the bearing, r is the radius of the

journal and R ' is the magnitude of the reaction in the joint.

In conclusion, a friction couple opposed to rotation trend, of moment ranging from


zero to a maximum value equal to the product between a friction coefficient in bearing µ ' ,
the radius of the journal r and the magnitude of reaction:
→ → →
M f ≤M f ,max = µ 'r R ' (6.21)
occurs in a bearing or joint (Figure 6.16) .

The friction coefficient µ ' , that is determined experimentally, depends on nature of


the surfaces in contact and the dimensions of the bearing as well as on the distribution of

the basic reactions on the contact surface. The magnitude of reaction R ' is expressed as:

(R ) + (R )

• R' = ' 2
x
' 2
y - for the cylindrical joint;

(R ) + (R ) + (R )

• R' = ' 2
x
' 2
y
' 2
z - for the spherical joint.
Figures T 6.15 a, b show examples of sliding and rolling bearings.

6.4.6. Belt friction

Let us consider a wheel with a belt wrapped around a portion of it. If the wheel is
fixed and the belt tends to move or the belt is fixed and the wheel tends to move, then
friction occurs between the wheel and the belt (Figure T 6.17). Denoting by µ the friction
coefficient, with θ the angle of wrap and with T 1 and T 2 the tensions in the belt on both
sides of the wheel, then the condition of equilibrium is written:

e−µ θ ≤ 1 ≤ eµ θ
T
(6.22)
T2
Relation (6.20) is called Euler’s formula for belt friction.

88
Theoretical mechanics – theory and applications

Tendinţa de mişcare

θ
T1
O•
T2

Figure T 6.17

6.5. Solved problems

R6.1. Given the homogenous beam AB of weight G and length l, simply supported at B on
a plane inclined at the angle β with the horizontal and jointed at A (frictionless).
Direction of the beam is at angle α with the horizontal (Figure R 6.1.1). An inextensible
string passed over frictionless fixed pulley at D, joins point C of the beam with a weight Q
(AC = a, γ given). If the beam is solicited by a couple of forces of moment M, find:
a) Reactions at A and B;
b) Value of moment M for which the beam stays in equilibrium in the given
position if the support at B is missing.

NB β
B B

 β β
G, l C
γ 
M γ
 y
M C Sc = Q

D VA

G

α Q
⋅ α
z
A 
HA x

Figure R 6.1.1 Figure R 6.1.2

Solution: a) Release the beam from its constraints (joint at A, string at C and the
simple support at B) and enter the corresponding reactions (see Figure R 6.1.2).
→ →
Since S C = Q , the unknowns of the problem are R A ( H A , V A ), R B ( N B ) . The scalar
equations of equilibrium are:

∑ X i = H A + Q cos(γ − α ) − N B sin β = 0 (1)


∑ Yi = V A − G − Q sin(γ − α ) + N B cos β = 0 (2)

89
Theoretical mechanics – theory and applications

+ N B ⋅ l cos( β − α ) = 0
l
∑ M A = M − G ⋅ 2 cos α − Q ⋅ a sin γ (3)

From equations (1-3) we get the reactions:


 a 1G M
NB = Q ⋅ sin γ + cos α − 
cos( β − α )  l 2 l 
H A = N B sin β − Q cos( γ − α ) , V A = G + Q sin( γ − α ) − N B cos β

b) The absence of the support at B leads to:


l
NB = 0 ⇒ M = G ⋅ cos α + Q ⋅ a sin γ .
2

R6.2. The bent beam OABC, of negligible weight, is fixed at O and solicited by:
- a uniformly distributed load, of intensity q (N / m) , located in the plane yOz ;
- a distributed load that decreases linearly from p (N / m) to 0 , located in plane
parallel to yOz ;
- a point load F (N), located in a plane parallel to xOz inclined at angle α with
Oz.
Dimensions of the beam are those in Figure R 6.2.1; determine the reactions at the
fixed support at O.
z 
Mz Q
z  3l
q
Rz Ry
A
O D A
y O y
2l 2l Mx My
l Rx F l
x B x B 

l α F l α F
s
2l
3l 
P
E

C  C′
p
C

Figure R 6.2.1. Figure R 6.2.2.

Solution: Forces uniformly distributed render a system of parallel forces and are
replaced with their resultant Q = 2 q l, with the point of application at the middle of the
segment AD (Figure R 6.2.2). Forces linearly distributed are also making a system of
3 pl
parallel forces and are replaced with the resultant P = ∫0 p ( s ) ds =
3l
(area of the
2
triangle BCC’), with the point of application at the center of the parallel forces located on
BC at the distance:
∫03 l s ⋅ p( s) ds
BE = 3 l = 2l .
∫0 p( s) ds
90
Theoretical mechanics – theory and applications

( )
Releasing the body from its constraint (fixed support at O) and entering the
components of the fixed support reactions R x , R y , R z and of the moment at the
( )
support M x , M y , M z , six scalar equations of equilibrium may be written as follows:
∑ X i = R x − F sin α = 0 ∑ M i x = M x − Q ⋅ 3l − F cos α ⋅ 4 l − P ⋅ 2l = 0
∑Y i = R y − P = 0 ∑ M i y = M y + F cos α ⋅ l = 0
∑ Z i = R z − Q − F cos α = 0 ∑ M i z = M z + F sin α ⋅ 4l − P ⋅ 2l = 0 .
By solving this system we get:

R x = F sin α M x = 6 q l 2 + 3 pl 2 + 4 F l cos α
3 pl
R y= M y = − F l cos α
2
R z = 2q l + F cos α M z = 3 p l 2 − 4 F l sin α .

R6.3. The rectangular plate ABCD is supported on a horizontal plane at A and on an


inclined plane at B (Figure R 6.3.1). Knowing the length of the rectangle sides AB = b and
AD = a, the sliding coefficient of friction µ at A and B, and the inclination angle α ,
determine the position of equilibrium given through the angle θ . Particularize the found
π a
result for θ = and → 0 (case of the beam supported on two perpendicular walls).
2 b
C
C b
D
D
a
O

 NB
B 
NA G 
A TB
θ α
β α
B
θ

A TA

Figure R 6.3.1 Figure R 6.3.2

Solution: Release the body from its constraints (friction supports at A and B) and
→ →
enter the matching reactions R A ( N A , T A ) , R B ( N B , TB ) (Figure R 6.3.2). The equations of
equilibrium and the conditions of equilibrium (at limit) associated to the friction at A and
B are:
∑ X i = T A + TB cos( π − α ) − N B sin ( π − α ) = 0
∑ Y i = N A − G + TB sin ( π − α ) + N B cos( π − α ) = 0
a2 + b2
∑ M i A = −G ⋅ 2 cos( β + θ ) + T B ⋅ b sin ( α + θ ) − N B ⋅ b cos( α + θ ) = 0
T A= µ N A , T B = µ N B .

91
Theoretical mechanics – theory and applications

Solving this system we get:


sin α + µ cos α µ
NA = G , N B= G , T A= µ N ,T B= µ N B ,
( µ 2 + 1) sin α ( µ 2 + 1) sin α A


1+ (ctg α − µ )
µ 2+ 1
tg θ = .

a
+ ( µ ctgα + 1 )
b µ 2+ 1
a π
Note: If → 0, α = (case of a beam supported with friction between two perpendicular
b 2
1− µ 2
walls), then the equilibrium is achieved for tg θ = .

R6.4. The system in Figure R 6.4.1 is made of a homogenous disk of weight G and radius
R, supported on a horizontal beam of negligible weight jointed at A and supported at B,
and a body of weight P situated in a smooth plane inclined at an angle α with the
horizontal direction. Knowing that AC = CB = l, R = l / 3, G = 1000 N, α = 300 and that
at C there is both sliding and rolling friction of coefficients µ = 0.1 and s = 0.15 R,
respectively, it is requested:
a) Values of the force P per equilibrium;
b) Reactions at A and B for P = Pmax .
C Tendinţa de
 alunecare Tendinţa de
P rostogolire


P
G, R µ, s 
 TA
C
A Mr
α y 
B O α NA

Figure R 6.4.1 Figure R 6.4.2


→ →
Solution: Under the action of the forces G and F , the half-disk may slide and roll,
respectively, in the sense shown in Figure R 6.4.2. or in the opposite sense. Releasing the
body from the friction support at A and entering the matching reactions (normal
→ → →
reaction N A , sliding force of friction T A and the rolling friction moment M r ), we write
the equations and conditions of equilibrium as follow:
∑ X i = P + T A cos α − N A sin α = 0 (1)
∑ Y i = −G + T A sin α + N A cos α = 0 (2)
∑ M i A = − M r + G ⋅ R sin α − P ⋅ R cos α = 0 (3)
No-slip condition: T A ≤ µ N A (4)
No-rolling condition: M r ≤ s N A (5)

92
Theoretical mechanics – theory and applications

From (1 – 5) we find that:


 s 
 tg α − µ tg α − 
R
P min = max  G, G (6)
 µ tg α + 1 s
tg α + 1 
 R 
Studying the case of opposed tendency of sliding and rolling respectively, we find
a maximum value for the force P in case of equilibrium:
 s 
 tg α + µ tg α + 
R
P max = min  G, G (7)
 − µ tg α + 1 − s tg α + 1 
 R 
For the half-disk to stay in equilibrium in the position in Figure R 6.4.2.it is
required that Pmin ≤ P ≤ P max .
Particular case: In the case of the particular values proposed for the parameters of
the problem, we find that:
tg α − µ tg α + µ
G = 33,77 N G = 87,79 N
µ tg α + 1 − µ tg α + 1
s s
tg α − tg α +
R R
G = 45,07 N G = 70,18 N .
s s
tg α + 1 − tg α + 1
R R
The following situations are possible:
a) P ∈ [0 ; 33,77 N ) - the body slips towards B and rolls in trigonometric
direction;
b) P ∈ [33,77 N ; 45,07 N ) - the body rolls in the trigonometric direction
without slippage;
c) P ∈ [45,07 N ; 70,18 N ] - the body is in equilibrium;
d) P ∈ (70,18 N ; 87,79 N ] - the body rolls clockwise without slippage;
e) P ∈ (87,79 N ; ∞) - the body slips towards C and rolls clockwise.
6.5. The homogenous plate in Figure R.6.5.1 has a joint at O characterized by the
coefficient of friction µ 0 and the journal radius r0 . Determine the values of the coefficient
of friction for the plate to stay in equilibrium in the position shown in the figure.
y
a
C2

a 
 V0
3a G2 Mf

 C1
a
H0
O x
a 
4a G1
4a
3

Figure R 6.5.1 Figure R 6.5.2

93
Theoretical mechanics – theory and applications

Solution: We release the body from its constraint (joint with friction at O) and
enter the matching reactions (components H O , VO of the reaction in the joint and the
friction moment in the joint M f ). To determine the center of mass of the plate we divide it

into two parts (the triangular surface and the square surface) and introduce the weights G 1

and G 2 into their centers of mass C 1, C 2 (Figure R 6.5.2). We further consider the
tendency of rotation of the plate as clockwise. The equations of equilibrium and the
condition of equilibrium associated to this type of friction are:
∑ X i= −H O= 0 (1)
∑ Y i = V O − G 1− G 2 = 0 (2)
3a a
∑ M iO = M f + G 2⋅
2
− G 1⋅ = 0
3
(3)

M f ≤µ 0 r 0 H O+ V O
2 2
(4)

Considering that the superficial density is ρ , we may write:

G 1= M 1 g = A 1 ρ g = 6 ρ a 2 g , G 2 = M 2 g = A 2 ρ g = ρ a 2 g (5)
a
From (1 – 5), we get µ 0 ≥ .
14 r 0
Note: The study of tendency of rotation in the trigonometric direction shows that the plate
remains in equilibrium irrespective of values of the coefficient of friction µ 0 .

R6.6. A cable passed over two drums holds attached two weights P and Q at its ends
(Figure R 6.6.1). Knowing the coefficients of friction between the cable and the two
drums, µ 1 and µ 2 respectively, find the ratio of the weights P and Q for equilibrium.

 
T −T
 
  P Q
P Q

Figure R 6.6.1 Figure R 6.6.2

Solution: We cut the cable in the portion between the drums and introduce the tension in
cable T (Figure R 6.6.2). The conditions that the two pieces of cable do not move on the
two drums are written as:

 3π  Q  3π   3π  T  3π 
exp − µ 1⋅  ≤ ≤ exp µ 1⋅  , exp − µ 2 ⋅  ≤ ≤ exp µ 2 ⋅ 
 2  T  2   2  P  2 

94
Theoretical mechanics – theory and applications

Performing the member by member multiplication of the two double inequalities


we get the values the ratio Q / P can take at equilibrium:

 3π  Q  3π 
exp − ( µ 1+ µ 2 ) ⋅  ≤ ≤ exp ( µ 1+ µ 2 ) ⋅ .
 2  P  2 

6.6. Problems proposed

6.6.1. Conventional tests

TC 6.1) Let the homogenous bar AB be of weight G and length l, simply supported
l
without friction at points A and D on a half-circle surface of radius R = (Figure TC
2
6.1.1). Determine the reactions at A and D and the position of equilibrium given by the
angle θ .

P
a a

B p C
D
O D
θ
R a

A 2a B 300

Figure TC 6.1.1 Figure TC 6.2.1

TC 6.2) Let the bar be as in Figure TC 6.2.1, jointed at A and simply supported at B. It is
acted upon on the AC side by the linearly distributed force of maximum intensity p (N/m)
and the concentrated force P (N) at point E. Dimensions of the bar are given in the figure,
find the reactions in the joint A and the support B.

l0
b
k
G G
r

µ r0 , µ 0 2r
α

Figure TC 6.3.1 Figure TC 6.4.1

95
Theoretical mechanics – theory and applications

TC 6.3) ) The homogeneous rectangular parallelepiped in Figure TC 6.3.1., with base side
a and height b is supported on an inclined plane whose angle α can be modified, the
coefficient of friction between the parallelepiped and the plane is µ . Determine angle α
for equilibrium.

TC 6.4) A bent bar, of negligible weight, is attached to an elastic spring of spring constant
k and initial length l 0 at one end and fitted with a ball of weight G at the other end.
(Figure TC 6.4.1). The bar is jointed with friction at point O. Knowing the position of
equilibrium given by angle ϕ , joint radius r 0 and the friction coefficient in the joint µ 0 ,
determine the values of the weight G at equilibrium.

6.6.2. Multiple-choice tests

TG 6.1) The homogenous plate in Figure TG 6.1 is composed of a portion bordered by a


half-circle of radius R continued with a portion of isosceles triangle shape with the base 2R
and height h. Determine the relation between the height h and the radius R so that the plate
stays in equilibrium on a horizontal plane for any of its inclination.
a) R = 2 h ; b) R = h ; c) h = 2 R ; d) h = 2 R .
a  
P Q

P a b
h
O
Mg B
R h A
µ, r

Figure TG 6.1 Figure TG 6.2 Figure TG 6.3

TG 6.2) A box of mass M = 75 kg and side a = 0.6 m rests on a rough surface (Figure TG
6.2). The contact between the box and the surface is characterized by the sliding
coefficient of friction µ = 0,2 . What is the maximum value of force P acting on the box
and what is the maximum height h for applying this force that will not allow the box to
either slip on the floor or to tip? Consider g = 9,81 m / s 2 .
a) Pmax = 147,15 N , hmax = 1,5 m ; b) Pmax = 125,28 N , hmax = 1,2 m ;
c) Pmax = 140,31 N , hmax = 2 m ; d) Pmax = 240,75 N , hmax = 1,8 m .

TG 6.3) The horizontal bar AB in Figure TG 6.3 has a joint at O characterized by the
coefficient of friction µ and the radius of the journal r. It is acted upon by the vertical
forces P and Q, found at the distances a and b, respectively, from the joint O. Specify the
range of values the ratio P / Q can take in order to achieve equilibrium in the position
shown in the figure.

96
Theoretical mechanics – theory and applications

P b − µ r b + µ r  P a + µ r a − µ r 
∈  ; b) ∈ 
Q  b − µ r b + µ r 
a) , , ;
Q a + µ r a − µ r
P r − µ b r + µ a
; d) ∈ (0, ∞) .
P
c) ∈ , 
Q r + µ b r − µ a Q

6.7. Indications and answers


TC 6.1) Release the bar from its constraints (frictionless supports in A and D) and enter
→ →
the normal reactions N A and N D (Figure TC 6.1.2). The equations of equilibrium are:
π −θ
∑ X i = N A cos( π − θ ) − N D sin 2 = 0
π −θ
∑ Y i = N A sin ( π − θ ) − G + N D cos 2 = 0
θ l π −θ
∑ M iA = N D ⋅ 2 R sin 2 − G ⋅ 2 cos 2 = 0 .
Solving this system we find:
θ
1 + 33 G cos 2 G
θ = 2 arcsin , NA = = 0,447G , N D = .
8 2 cos θ 2

P
a a

C D

y  a
ND B pa

O ⋅ π − θ/ 2 B
 θ D x NB
NA  C 4a / 3 300
G
A
HA
VA

Figure TC 6.1.2 Figure TC 6.2.2

TC 6.2) The linearly distributed force is replaced with the resultant R = p a placed on the
bar AC at the distance 2a / 3 from the point C (Figure TC 6.2.2). Enter the reactions
→ →
R A (H A , VA ) and R B (N B ) respectively, and write the equations of equilibrium:
∑ X i = H A − N B sin 30 0 + pa = 0
∑ Y i = V A + N B cos 30 0 − P = 0

97
Theoretical mechanics – theory and applications

4a
∑ M iA = N B sin 30 0 ⋅ a + N B cos 30 0 ⋅ a − P ⋅ 2a − pa ⋅ 3
= 0.

The unknowns of this system are H A , V A and N B .

TC 6.3) The normal reaction from the plane moves its point of application with increasing
plane angle. At the limit of the equilibrium, which may get lost by tipping over, the
reaction is applied at the extreme position A of the base (Figure TC 6.3.2). The reduction
of the joint forces at the center O of the se requires introduction of the moment of rolling
friction M r . The equations and conditions of equilibrium are written:

∑ X i = T − G sin α = 0
∑ Y i = N − G cos α = 0
b
∑ M iO = M r − G ⋅ 2 sin α = 0
T≤µN , M r≤sN.
The coefficient of rolling s is the maximum displacement the normal reaction N can have
a a
i.e. s = . Solving the set of conditions from above, we get α ≤ min(arctg µ , arctg ) .
2 b
y

l0
C

r G
G

x Mf
O T x
Mr
α ϕ Fel = k ⋅ 2r sin ϕ
N
2r

Figure TC 6.3.2 Figure TC 6.4.2

TC 6.4) The elastic force developing in the spring is Fel = k ⋅ 2 r sin ϕ and the joint with

friction at O is replaced with the reaction R '
O (H O , VO ) and the moment of friction M f
(Figure TC 6.4.2). Considering the tendency of rotation of the bar in the clockwise
direction, the equations of equilibrium are:

∑ X i= HO = 0
∑ Y i = V O − G + F el = 0
∑ M iO = M f + Fel ⋅ 2 r cos ϕ − G ⋅ r sin ϕ = 0
M f ≤ µ0 r0 H O+ V O
2 2

98
Theoretical mechanics – theory and applications

Considering the opposing tendency of motion, it is found that:


2 r k sin ϕ (2 r cos ϕ + µ 0 r 0 ) 2 r k sin ϕ (2 r cos ϕ − µ 0 r 0 )
≤G≤ .
r sin ϕ + µ 0 r 0 r sin ϕ − µ 0 r 0

TG 6.1) It is required that the resulting moment of the weight forces is zero:
π
2 sin 2 h
G 1⋅ R = G 2⋅ .
3 π 3
2
Correct answer: c)

TG 6.2) The equations of de equilibrium are written:


a
N = G = M g , P = T ≤ µ N = Pmax , P ⋅ hmax = Mg ⋅ .
2
Correct answer: b).

TG 6.3) For the tendency of rotation in the trigonometric direction, the equations of
equilibrium and the condition of equilibrium are written:
∑ X i= HO = 0
∑Y i = V O − P − Q = 0
∑MiO = Q ⋅b − P ⋅ a + M f =0

Mf ≤µr H O+ V O
2 2
.
Correct answer: a).

99
Theoretical mechanics – theory and applications

7. Equilibrium of the systems of material points and of solid rigids


7.1. Conditions of equilibrium of a system of forces acting upon a system of material
points

A set of n material points, somehow connected to each other are forming a system.
The forces acting on the material point A i , i = 1, n , are of two kinds:

- external forces, whose resultant we shall denote by F i (Figure T 7.1);

- internal forces , denoted F i j , j = 1, n, j ≠ i .
According to the principle of action and reaction the internal forces are equal two by
two and directly opposite:
→ → →
F i j + F j i = 0 , i, j = 1, n, j ≠ i (7.1)
In addition, the moment about an arbitrary point O of any pair of internal forces
→ →
F i j , F j i is null:
→ → → → → → →  →  → →  →
r i× F i j + r j× F j i = r i× F i j + r j× − F i j  =  r i − r j  × F i j = (7.2)
   
   
→ → → → →
= A i A j × F i j = 0 , because A i A j // F i j .
  Aj

 
Ai Fij Fji
   

A1

A2
Fi Fj F −F

   
ri rj r1 r2

•O
Figure T 7.1 Figure T 7.2

By definition, a system of material points is in equilibrium if every point in the


system is in equilibrium and vice versa.
The necessary and sufficient condition that a point A i , i = 1, n , in the system is in
equilibrium:
→ n → →
Fi+ ∑Fij = 0 , i = 1, n (7.3)
j =1
j ≠i

100
Theoretical mechanics – theory and applications

Summating equations (7.3) we find that:


n → n n → →
∑ Fi + ∑∑Fij = 0 (7.4)
i =1 i =1 j =1
j ≠i

Performing the scalar multiplication of relations (7.3) by vectors r i , i = 1, n , to the
left, we get the equations:
→ → n → → →
r i× F i + ∑ r i × F i j = 0 , i = 1, n (7.5)
j =1
j ≠i
which, by addition, lead to:
n → → n n → → →
∑ r i× F i + ∑∑ r i× F i j = 0 (7.6)
i =1 i =1 j =1
j ≠i
n n → → n n → → →
As per relations (7.1 – 7.2) we have that ∑∑ Fij = 0 , ∑∑ r i× F i j = 0 ,
i =1 j =1 i =1 j =1
j ≠i j ≠i
so that , from (7.4) and (7.6), we obtain the following conditions necessary for equilibrium
for a system of material points:
n → → n → → →
∑Fi = 0 , ∑ r i× F i = 0 (7.7)
i =1 i =1

But these conditions are not sufficient for the equilibrium of a system of material
points because they do not ensure that each point of the system is in equilibrium. Thus,
considering a system of two material points, A 1 and A 2 acted upon by the collinear forces
→ →
F and - F (Figure T 7.2), although the relations (7.7) are verified, the points will not be in
equilibrium but they will get closer to each other.
For a rigid solid, the conditions (7.7) are necessary and sufficient conditions of
equilibrium.

7.2. Theorems and methods for solving problems of statics of systems

Conditions (7.7) represent the mathematical expression of the theorem of


solidification whose statement is:
Theorem of solidification: A system of deformable points (or rigids), which is in
equilibrium under the action of an arbitrary system of forces, continues to remain at rest if
it is reinforced by the introduction of additional internal connections.

Should only a part of the equations (7.3) and (7.5) are considered, namely those
expressing the conditions of equilibrium for the points forming a system of points isolated

101
Theoretical mechanics – theory and applications

from the initial system of points, proceeding as above, conditions similar to those
expressed by (7.7) are obtained, noting that the sums are considered solely for the points of
the subsystem. The theorem of the equilibrium of the parts is obtained, stated as follows:

Theorem of the equilibrium of the parts: If a deformable system of points (or


rigids) is in equilibrium under the action of system of forces, then any part of the system
considered as rigid is also in equilibrium under the action of forces corresponding to that
part.

Using this theorem one can see that to in order to solve a problem of equilibrium
of body systems the equilibrium of the individual body equilibrium may be considered. So
we get to the free-body method.
Each free body will be acted upon by the given external forces (active), by the
external forces of connection (unknown) from the connection of the body to the bodies
from outside the body and by the internal connection forces (unknown) from the
connections of this body to the other bodies in the system.
The types of connections (internal or external) are the same met in the case of the
rigid with connections and their treatment is the same. As concerns the internal connection
forces, the principle of action and reaction must be carefully complied with, meaning that
if there is an action on a body, there will be a an equal and directly opposite reaction in the
body that comes in contact with it. In solving problems of equilibrium of body systems, the
parameters fixing the equilibrium position as well as the reaction components occur as
unknowns. In order to remove some reactions which are not of interest, the theorem of
equilibrium of the parts or even the theorem of solidification may be applied, that is the
equations of equilibrium for just a part of the system considered as rigidised or for the
whole system considered as rigid are written.

7.3. Solved problems

R 7.1) Given the system of bars in equilibrium shown in Figure R 7.1.1. Knowing
that l = 1 m, q = 2 kN / m, P = 10 kN and M = 20 kN m, find the reactions in the
fixed support A, simple support B and joint C.

q
D
l
C

M 
P
3l •

A
B
2l 1,5l 1,5l

Figure R 7.1.1

102
Theoretical mechanics – theory and applications

Solution: Method of free-body is applied, considering each body separately, acted


upon by the external forces, forces of connection to the outside of the system and the
forces of connection between the bodies composing the system. The uniformly distributed
force q is replaced with an equivalent concentrated force Q = 4 l q, applied at the middle of
the bar AD and having the same direction and sense with the force q. The three equations
of equilibrium corresponding to the plane case are written for each body.

Body 1 (figure R 7.1.2)

∑ X i = −H A + Q − HC = 0 (1)
∑Yi = V A − VC = 0 (2)
∑ M i A = M A − M − Q ⋅ 2l − VC ⋅ 2 l + H C ⋅ 3 l = 0 (3)

D
HC
l M C
C HC
2l P
VC
l VC
Q
3l •

2l NB
450
MA B
1,5l 1,5l
HA
A
VA

Figure R 7.1.2 Figure R 7.1.3

Body 2 (figure R 7.1.3)

∑ X i = H C − P sin 450 =0 (4)


∑Yi = VC + N B − P cos 450 = 0 (5)
3 2l
∑ M iC = N B ⋅3l − P ⋅
2
=0 (6)

From equations (1 – 6) it is obtained:

2
At B: N B = P = 5 2 kN
2
At C: H C = P sin 45 0 = 5 2 kN ; VC = P cos 45 0 − N B = 0
( )
H A = − H C + Q = 8 − 5 2 kN ; V A = VC = 0 ;
At A:
M A = M + 2 l Q + 2 l VC − 3 l H C = 28 − 15 2 kN ⋅ m

103
Theoretical mechanics – theory and applications

R 7.2) The bar AB, of weight G and length 4 l, is supported without friction at points A
and C (AC = 3 l). The end A is connected to an inextensible horizontal string AD and the
bar forms an angle α with the horizontal direction. On the bar, at point E (AE = l) there is
a sphere of radius R and weight P attached to the string OF, parallel to the bar AB (Figure
R 7.2.1). Knowing that the system is in equilibrium, find the tensions in the strings AD and
OF.

C
O
E
α D
A

Figure R 7.2.1

Solution: Single out the two bodies (the bar and the sphere) and write the
equations of equilibrium:
Bar AB (Figure R 7.2.2)

∑ X i = T AD + N E sin α − N C sin α = 0 (1)


∑Yi = N A − G − N E cosα + N C cos α = 0 (2)
∑ M i A = −T AD ⋅ l − G ⋅ 2 l cos α + N C ⋅ 3 l = 0 (3)

Sphere of center O (Figure R 7.2.3)

∑ X i = TOF − P sin α = 0 (4)


∑Yi = N E − P cos α = 0 (5)

y  y
NC
x

l 
x TOF
⋅ B
l
 C
NA l O
E  α
l G

α D NE
 
A TAD NE 
P

Figure R 7.2.2 Figure R 7.2.3

From (1 – 5) results:

104
Theoretical mechanics – theory and applications

N E = P cos α ; N C =
(2G − P sin α ) cos α
; NA =
3 P cos 2 α + G 3 − sin α − 2 cos 2 α ( )
3 − sin α 3 − sin α
( 2G − 3P) sin α cos α
TOF = P sin α ; T AD =
3 − sin α
R 7.3) Let the rigid bodies system be in Figure R 7.3.1. Bar OA, of length l and negligible
weight, is fixed at O and joined at A with the hoist of radii R1 and R 2 respectively, and
weight Q. The radius of the joint journal is ρ and the friction coefficient is µ . Over the big
sheave of the hoist a cord is wrapped having attached a body of weight P at one end. To
the small sheave of the hoist it is attached, by a cord as well, the homogenous flat plate of
weight G shown in the figure.
Given the values for l , R 1 , R 2 , a, ρ , µ , Q, G , it is requested:
a) The maximum value of the force P at equilibrium;
b) Reactions at O and A for the maximum value of force P.

R 1 , R 2 (R 1 < R 2 )
2

O l
A
1
 2a
Q C B
 a  3
4 P
G

Figure R 7.3.1
Solution: a) Single out the bodies of the system, introduce the external and the
connection forces and write the equations and conditions of equilibrium in the limit. The
trend of motion of the system corresponding to the maximum value of P is the one in
which body 4 is descending.
Body 1 (Figure R 7.3.2)
∑ X i = H A − HO = 0 (1)
∑Yi = − V A + VO = 0 (2)
∑ M iO = M f − M O + VA ⋅ l = 0 
(3)
− VA

y

 
 − Mf
V0   − HA
M0 Mf A
HA
l
R2
 0 A x R1
H0

VA  
T2 Q T1

Figure R 7.3.2 Figure R 7.3.3

105
Theoretical mechanics – theory and applications

Body 2 (Figure R 7.3.3)

∑ X i = −H A = 0 (4)
∑Yi = V A − T1 − T2 − Q = 0 (5)
∑ M i A = T2 ⋅ R 2 − T1 ⋅ R 1 − M f =0 (6)

Mf = µ ρ H 2A + V 2
A
(7)
Body 3 (Figure R 7.3.4)
∑ Yi = T1 + T3 − G = 0 (8)
∑ M i B = G ⋅ b − T1 ⋅ 2a = 0 (9)
b is the abscissa of the center of mass, that can be determined depending on a :
13 − 3 π
b= a
3( 6 − π ) y y

− T1 
2a − T3 
b − T2
x C
B

x
a

Pmax

G

Figure R 7.3.4 Figure R 7.3.5

Body 4 (Figure R 7.3.5)

∑Yi = T2 − Pmax = 0 (10)

( R 1+ µ ρ ) b
G + µ ρQ
2a
From equations (1 – 10) we get Pmax = .
R 2− µ ρ

b) Reactions at joint A and embedment O are determined from the equations (1 – 6). Their
components are:

b
At A: H A = 0 ; V A = Pmax + G + Q ; M f = µ ρ VA
2a

G + Q ; M O = ( µ ρ + l ) VO
b
At O: H O = 0 ; VO = Pmax +
2a

106
Theoretical mechanics – theory and applications

R 7.4) Let the system of bodies be in Figure P 7.4.1 composed of:


- The bent bar ABC ( AB = a, BO = l 1 , CO = l 2 ), of negligible weigh, joined at
O (frictionless) and acted upon at C by the vertical force F (unknown);
- The hoist of weight G and radii r, R. Joint at O 1 is frictionless ;
- The weight Q attached to the small sheave of the hoist by means of a cord
(frictionless).
Between the cord attached at A and D and the big sheave of the hoist there is a friction
coefficient µ . Requested:
a) The range of values of the force F in order to get the equilibrium for the position
shown in the figure;
b) Reactions at the joints O and O 1 for F = Fmin .


r, R , G A
B O C


F
O1
µ

 D
Q

Figure R 7.4.1

Solution: a) The method of free-body is applied.

Body 1 (Figure R 7.4.2)

∑ X i = H O − T 1= 0 (1)
∑Y i = V O − F = 0 (2)
∑ M iO = T 1⋅ a − F ⋅ l 2 = 0 (3)
T1
T1
V0
A 1
R

a V0
H0 O1 r
H0 1
O Q
B C G
l1 l2 T2
F

Figure R 7.4.2 Figure R 7.4.3

107
Theoretical mechanics – theory and applications

Body 2 (Figure R 7.4.3)


∑ X i = H O − T 1− T 2 = 0
1
(4)
∑ Y i = VO − G − Q = 0
1
(5)
∑ M O = Q ⋅ r + T 2⋅ R − T 1⋅ r = 0
1
(6)
In addition, we have Euler’s relation, too:
T 2 exp(− µ π ) ≤ T 1 ≤ T 2 exp( µ π ) (7)
r a exp(µ π )
From (3), (6) and (7) we get F ≥ Fmin = Q .
R l 2 exp(µ π ) − 1
l r exp( µ π )
b) At O: H O = T 2 = Fmin =Q
2
, VO = Fmin
a R exp( µ π ) − 1
r exp( µ π )
At O 1 : H O 1 = T 2 + T 1= 2 Q , VO ! = G + Q
R exp( µ π ) − 1
R7.5. Let us consider the system of rigid bodies in figure R 7.5.1 composed of:
- The hoist of weight G0 , radii r 0 , R 0 and moment of inertia about the axis of
symmetry equal to J 0 , resting on a horizontal plane characterized by the
rolling friction coefficient s and the sliding friction coefficient µ (unknown);
- The fixed pulley block of weight G 1 and radius R 1 , joined at O 1 with the
vertical bar AO 1 ;
- The bar AO 1 , of length l and negligible weight, embedded at A ;
- The mobile pulley block, of weight G 2 and radius R 2 ;
- The weight G 3 ;
The bodies are connected to each other by inextensible cords and maintained in
equilibrium as shown in the figure due to the couple of moment M applied to the hoist.
Determine the coefficients of friction µ and s, so that the equilibrium is not broken for the
lowering sense of the weight G 3 .


J 0 , r0 , R 0 , G 0
R 1 , G1

O1 O M


R2, G2
I ( µ, s )
l
O2


G3
A

Figure R 7.5.1

108
Theoretical mechanics – theory and applications

Solution: Body isolation method is applied.

The hoist (Figure R 7.5.2)


∑ X i = TIi − S1 = 0 (1)
∑Yi = N I − G0 = 0 (2)
∑ M i I = M + M r − S1 (R 0 + r 0 ) = 0 (3)
TI ≤ µ N I (4)
Mr ≤ s N I (5)

VO 
1
− S1
S1

H0
r0 M 1
O O1

R0 G0
 y
NI S2 
TI G1
Mr
x

Figure R 7.5.2 Figure R 7.5.3

Fixed pulley block (Figure R 7.5.3)

∑ X i = S1 − H O = 0 1
(6)
∑Yi = VO − G 1 − S 2 = 0
1
(7)
∑ M O = S 1 R 1 − S 2 R 1= 0
1
(8)

Mobile pulley block (Figure R 7.5.4)

∑Yi = S 2 + S 3− S 4 − G 2 = 0 (9)
∑ M O = S 3 R 2− S 2 R 2= 0
2
(10)
  O1
S3 − S2  
y − H0
1
− S4

− V0
1

O2


S4  x
HA
 A 
G2   G3
MA VA

Figure R 7.5.4 Figure R 7.5.5 Figure R 7.5.6

109
Theoretical mechanics – theory and applications

The bar (Figure R 7.5.5)


∑ X i = HO − H A = 0
1
(11)
∑Yi = V A − VO = 0 1
(12)
∑M i A = HO ⋅l − M A = 0
1
(13)

The weight G 3 (Figure R 7.5.6)


∑Yi = S 4 − G 3= 0 (14)
From relations (1 – 14) it results:
G 2+ G 3 G2 +G3
S 1= S 2 = S 3 = T I = H A = H O 1 = , S 4= N I = G 0, M A = l
2 2
2 G 1+ G 2 + G 3 G 2+ G 3
V A = VO 1 = , Mr = (R 0 + r 0 ) − M
2 2
G + G 3 not (G 2 + G 3 )(R 0 + r 0 ) − 2 M not
µ ≥ 2 = µ min , s ≥ = s min
2G 0 2G 0

7.4. Problems proposed

7.4.1. Conventional tests


TC 7.1) Let us consider the joined bars system in Figure TC 7.1.1. Given the forces P 1
→ → →
and P 2 , couples M 1 and M 2 , angle α as well as the distance a, determine reactions in the
supports A and B, joints C and D and embedment E.
 
P1 P2 

A M1 C α M2
B D E

2a 3,5a 2,5a 2a 2,5a 3,5a

Figure TC 7.1.1

TC 7.2) The system in Figure TC 7.2.1 is composed of a homogenous disk of weight G


and radius R that rests on a horizontal bar of negligible weight that has a joint at A and is
simply supported at B, and a body of weight P found on a smooth plane inclined at an
angle α with the horizontal direction. Knowing that AC = CB = l, R = l / 3, G = 1000 N,
α = 30 0 and that at C there is both sliding and rolling friction of coefficients µ = 0,1
and s = 0,15 R respectively, then find:
a) Values of force P for equilibrium;
b) Reactions at A and B for P = P max .

110
Theoretical mechanics – theory and applications

 150
G, R 200KN
A
A B F = 50KN

C (µ, s ) P
100KN
l l α
B

Figure TC 7.2.1 Figure TC 7.3.1

TC 7.3) Two stone blocks of weights G 1= 200 kN and G 2 = 100 kN , respectively, are
supported one to each other and on two perpendicular walls (Figure TC 7.3.1). The
coefficient of friction between any of two surfaces of contact is µ = 0,2 . Is the force F =
50 kN sufficient to maintain the blocks in equilibrium? Consider that the force F is applied
at the center plane of the two blocks so that we may consider that the problem is coplanar.

TC 7.4) Given the system of bodies in Figure TC 7.4.1. The bent bar OBC
  ∧ 
 mOBC = 90 0  is of negligible weight and is embedded at its end at O. At C there is a
  
hoist of radii R 1 and R 2 respectively, and weight G attached by means of a joint. The joint
at C is with friction of coefficient µ 0 and the radius of the joint journal is ρ . Over the
sheave of radius R 1 it is wrapped a cord that at the other end is attached to the center O 2
of a disk of weight Q, simply supported at A on the bar OBC, the direction of the cord
being reversed by means of a small pulley block (at K) of negligible mass friction. There is
sliding friction of coefficient µ and rolling friction of coefficient s at the support A. Over
the sheave of radius R 2 another cord is wrapped that at the other and is attached to a rotor
of radius R and weight G 1 , driven by a driving moment M m . The lengths OA = l, AB =
BC = 2l are also given. Requested:
a) Determine the maximum value of the driving moment M m at equilibrium;
b) Reactions in joint C and embedment O.

R1 , R 2 , G (R1 < R 2 ) R , G1

O1
C(µ 0 , ρ)

Mm
r, Q

O2
K

O A(µ, s) B

Figure TC 7.4.1

111
Theoretical mechanics – theory and applications

TC 7.5) Let us consider the rigid solid system in Figure TC 7.5.1 composed of:
- The bar O 1 E , of length R and negligible weight;
- The body of weight P;
- The disk of radius R and weight G with the center at point C;
- The bar O 2 B , of length 3R and negligible weight, inclined at the angle α with the
horizontal direction.
There is sliding friction of coefficient µ 1 between the cord and the disk, as well as
sliding friction of coefficient µ and rolling friction of coefficient s between the disk and
π 1 R 3 5
the bar O 2 B . Knowing O 2 D = R, α = , µ = ,s = , µ 1= ln , determine:
6 2 3 4 3 π 2
a) The range of values of the weight P for equilibrium;
b) Reactions at the joints O 1 and O 2 respectively, and at the support B for the
maximum value of P.
R

E O1
µ1

C
G


O2
µ R
2R

Figure TC 7.5.1

7.4.2. Multiple-choice tests

TG 7.1) What force F is required to move the body of mass M 2 = 300 kg to the right? The
coefficient of friction between any two surfaces in contact is µ = 0,3 (Figure TG 7.1.1).
Consider g ≅ 10 m / s 2 .
a) F ≅ 2473 N ; b) F ≅ 1583 N ; c) F ≅ 1182 N ; d) F ≅ 985 N .
B

100Kg
F 2m
200Kg 30 0

450 A

Figure TG 7.1.1 Figure TG 7.2.1

112
Theoretical mechanics – theory and applications

TG 7.2) What is the minimum value of the friction coefficient µ between the bar AB and
the support surfaces for the bar AB to remain in equilibrium in the position shown in
Figure TG 7.2.1?
Given the weight of the bar G = 200 N and its length L = 3.3 m.
a) µ min = 0,322 ; b) µ min = 0,127 ; c) µ min = 0,275 ; d) µ min = 0,405 .

7.5. Indications and answers

TC 7.1) See Figure TC 7.1.2.

y y y
   
    P2 VD P2 VD
NA P1 NB VC  

− HC α α
M1 D − HC D
B C C
A  C 
 x 2a 2,5a HD x x
2a 2,5a HD
2a 3,5a 2,5a HC  
− VC
− VC

Figure TC 7.1.2

The equations of equilibrium for the three bodies are:


H C = 0, N A − P 1+ N B + VC = 0,− N A ⋅ 8 a + P 1⋅ 6a − M 1− N B ⋅ 2,5 a = 0
H D + P 2 cos α − H C = 0, V D − P 2 sin α − VC = 0, − VC ⋅ 4,5 a − P 2 sin α ⋅ 2,5 a = 0
H E − H D = 0, V E − V D = 0, − M E − M 2 + V D ⋅ 3,5 a = 0 .
The solution of this system is:
7 2 M 1 25
At A: N A = P 1− − P sin α ;
11 11 a 29 1
4 2 M 1 80
At B: N B = P 1− + P sin α ;
11 11 a 29 1
5
At C: H C = 0 , VC = − P 1sin α ;
9
4
At D: H D = − P 2 cos α , V D = P 2 sin α ;
9
4 14
At E: H E = − P 2 cos α , V E = P 2 sin α , M E = P a sin α − M 2 .
9 9 2

TC 7.2) See Figure TC 7.2.2. The trend of motion of the disk is from A to B.
y

 y
T1 T1 
N0 VA VB
Mr
 D HA T
G
α  x B
 A
T  x P l l
Mr
N

α N
Figure TC 7.2.2

113
Theoretical mechanics – theory and applications

The equations and conditions of equilibrium for the three bodies are:

T 1− T = 0, N − G = 0, M r − T 1⋅ 2 R = 0, T ≤ µ N , M r ≤ s N (1)
P sin α − T 1= 0 , N D − P cos α = 0 (2)
T − H A = 0, V A + V B − N = 0, − M r − N ⋅ l + V B ⋅ 2l = 0 (3)
 µ s 
From (1) and (2) we find that P ≤ Pmax = max G, G  = 200 N .
 sin α 2 R sin α 
For P = Pmax , the values of the reactions are:
G R
H A = Pmax sin α = 75 N , V A = − Pmax sin α = 475 N ,
2 l
G R
V B = + Pmax sin α = 525 N .
2 l

TC 7.3) See Figure TC 7.3.2 (block B tends to move to the left).

A
N1 G1
150

T1
T2 N2
N2 T2

F
G2
B
T3
N3

Figure TC 7.3.2

The scalar equations of equilibrium associated to the two bodies are:

 N − T cos 15 0 − N sin 15 0 = 0
1 2 2
Block A:  ;
 N 2 cos 15 − T 2 sin 15 0 − T 1− G 1= 0
0

T − F − T cos 15 0 + N sin 15 0 = 0


3 2 2
Block B:  ,
− N 2 cos 15 + T 2 sin 15 0 + N 3 − G 2 = 0
0

to which the conditions T 1 ≤ µ N 1 , T 2 ≤ µ N 2 , T 3 ≤ µ N 3 are added. F ≥ 174 N is


obtained, so that the force of 50 N is sufficient to maintain the equilibrium in the sense
under consideration. Identically, a study of the trend of motion of the block B to the right
shall be performed.

114
Theoretical mechanics – theory and applications

TC 7.4) See Figure TC 7.4.2.

VC V1
T2
T2
Mm
O2
T1
C H1
HC O1
Q
TA Mf

A
G1
Mr T1 G
NA
VC

HC
C
Mf

V0 NA

H0 TA
A
B
O Mr
M0

Figure TC 7.4.2
The equations and conditions of equilibrium for the four bodies are:

T 1− T A = 0, N A − Q = 0, M r − T 1 r = 0 , T A ≤ µ N A , M r≤ s N A ,
M f + T 1 R 1− T 2 R 2 = 0 , M f ≤ µ 0 ρ H C2 + V C2 ,
− H O + T A − H C = 0, V O − N A − VC = 0 ,
M O − M r − M f − N A ⋅ l + H C ⋅ 2l − VC ⋅ 3l = 0 ,
H 1− T 2 = 0 , V 1− G 1= 0 , M m − T 2 ⋅ R = 0 .

max
Solving this system in the limit for M m the inequations become equations.
Found:
s Mm  s
H O=
Q+ , V O = 1 +  Q + G ,
R R  R
 3s 
l  Q + 3 l G + (2 l + R 2 )
s Mm
M O =  s + R − R 1+ ,
 r R  R
Mm s s R2
H C= − , V C = Q + G , M f = − R1 Q + Mm .
R R r R
The driving moment results from a second degree equation obtained by removing
the other unknowns (for µ 0 = 0 ). Its value is:
s R R1
M m= Q.
r R2

115
Theoretical mechanics – theory and applications

TC 7.5) See figure TC 7.5.2.

T1
N B
T N2
R
C V2 M
T O2 2R
α
G
H2
T2 M
N
T1
V1
M1
E O1
α R H1
T1 P

Figure TC 7.5.2

a) Write the equations and conditions of equilibrium and obtain a system of 10


equations and 3 inequations with 12 unknowns: H 1, V 1, M 1, T 1, N , T , M , T 2 , V 2 , H 2 , N 2
and P. Consider the double motion tendency when writing the conditions of equilibrium.

T + T 1− (G + T 2 ) sin α = 0 , N − (G + T 2 ) cos α = 0 , − M + (T2 + T − T 1 ) ⋅ R = 0 ,


H 2 − T = 0 , N 2 + V 2 − N = 0 , M + N 2 ⋅ 3R − N ⋅ 2 R = 0 (1)
H 1− T 1 cos α = 0 , V 1− T 1sin α = 0 , M 1− T 1 R sin α = 0, T 1− P = 0 .
π  π 
− µ  −α  µ  −α 
2  2 
− µ N ≤ T ≤ µ N,− s N ≤ M ≤ s N, T 2 e ≤ T 1≤ T 2 e (2)

Select as parameters the forces P and T. From (1) we find:

T 1= ( G + P) sin α − T , N = (G + P) cos α , M = R [2T + P − ( G + T ) sin α ] ,

T 2 = P , H 2 = T , V 2 = [( G + P)( cos α − sin α ) + P + 2 T ]


1
(3)
3
N 2 = [( G + P)( 2 cos α + sin α ) − P − 2 T ], H 1 = [( G + P) sin α − T ] cos α ,
1
3
V 1= [( G + P) sin α − T ] sin α , M 1= R [( G + P) sin α − T ] sin α .

Replacing these values in (2) we deduce the inequations:

P G P G
T+ + ≥0 , T − − ≤0,
4 4 4 4

116
Theoretical mechanics – theory and applications

5 P 3G 3P 5G
T+ − ≥0 , T + − ≤0 (4)
16 16 16 16
P G G
T− − ≤0 , T +2P− ≥0.
10 2 2
This system is resolved graphically in the system of coordinates (P, T). The
minimum value, Pmin , is obtained from the intersection of the lines of equation
P G G G
T − − = 0 and T + 2 P − = 0 . It equals to Pmin = . Corresponding to it we also
4 4 2 9
5G
have Tmin = . In addition, Pmax = ∞ .
18
Loss of the equilibrium can only be achieved by sliding of the cord over the disk
and lowering of the disk by frictionless rolling along the line O 2 B .
G 5G
b) For P = and T = the reactions are:
9 18

5 3 5 5 5 5 3 +1 10 3 − 1
H 1= G , V 1= G, M 1 = R G , H 2 = G, V 2 = G, N 2 = G.
36 36 36 18 27 27

TG 7.1) Isolate the two bodies (figure TG 7.1.2).

N1
• G1

T12
N12

N12
T12
F
30 0
•G 2

T2
N2

Figure TG 7.1.2

Writing the equations of equilibrium in the limit, we get the value:

(2 m 2 + m 1 ) g
F= ≅ 1182 N .
cos 30 0
sin 30 +
0
µ
Correct answer: c).

117
Theoretical mechanics – theory and applications

TG 7.2) See Figure TG 7.2.2.


TC
NC
B

G NA

450
A
TA

Figure TG 7.2.2

Write the equations of equilibrium on the horizontal and vertical directions and the
equation of moments about point A as well as the conditions of equilibrium in the
limit T A = µ N A , TC = µ N C . It is obtained µ min = 0,322 . Correct answer: a).

118
Theoretical mechanics – theory and applications

Kinematics
Kinematics studies the mechanical motion of bodies without taking into account
their mass and the forces acting upon them.
Fundamental notions of space and time are used in kinematics. The notion of
motion is a complex concept that encompasses several elements within its field:
- body in motion (also called movable);
- environment in which the motion takes place;
- the reference against which the motion is performed
If the reference is fixed, then the motion is called absolute and if the reference is
movable the motion is called relative.
To know the motion of a rigid means knowing the motion of any point of the rigid.
Hence, in what follows, the main kinematic concepts are set forth considering the material
point.

8. Kinematics of the absolute motion of the material point


8.1. Trajectory, velocity and acceleration

8.1.1. Trajectory

Motion of a material point M is known if you can specify its position relative to a

fixed point O at any instant, i.e. if the position vector r as a function of time is known:
→ →
r = r (t ) (8.1)

Function r is considered continuous, uniform, differentiable at least twice and
finished in magnitude (figure T 8.1).

M
S
 ( C)
r
M 0 ( t = 0) M ( t > 0)
O

Figure T 8.1 Figure T 8.2

119
Theoretical mechanics – theory and applications

Motion of a material point M is also known if curve (C) along which the point
moves (Figure T 8.2) and how the point moves along it is indicated, more specifically if
the law s = s (t), called hourly law of motion, is known.
The trajectory means the geometric place of the successive positions taken by the

point in its motion (or the geometric place of the extremities of the position vector r ).
Between trajectory and curve (C) along which the point is moving there is not always an
identity.

8.1.2. Velocity

Mobiles can travel the same distance in different time intervals or different
distances in the same time, so, in order to characterize the motion it is necessary to
introduce the notion of velocity.

By definition, the velocity (instantaneous) of a material point is given by the first


derivative of the position vector with respect to time:
→ •

d r not →
v = = r (8.2)
dt
Notes: i) In mechanics, the first order derivative of a scalar or vector function is marked
with a dot over the function name, the second order derivative by two points, etc.
→ → →
→ d r ∆r ∆r →
ii) It can be seen that v = = lim , where = v medie is the average velocity
d t ∆ t →0 ∆ t ∆t
→ →
over the time interval ∆ t , ∆ r being ‘’the increment’’ of the position vector r .
iii) Velocity unit in the SI is m ⋅ s −1 .

8.1.3. Acceleration

Is a vector quantity that characterizes the variation of velocity of a point in its


motion as direction, sense and magnitude.

By definition, acceleration (instantaneous) of a point is given by the first


derivative of the velocity vector with respect to time or the second derivative of the position
vector with respect to time:
→ • → ••
→ d v not → d 2 r →
a = = v = = r (8.3)
dt dt2
→ → →
→ d v ∆v ∆v →
Notes: iv) It may be seen that a = = lim , where = a medie is
d t ∆ t →0 ∆ t ∆t

120
Theoretical mechanics – theory and applications


average acceleration over the time interval ∆ t , ∆ v being the'' increment'' of the velocity
vector.
v) Acceleration unit in the SI is m ⋅ s −2 .

8.2. Analytical expressions of velocity and acceleration


in different systems of coordinates

8.2.1. Frenet System of coordinates (intrinsic, natural)

Frenet’s frame is a moving trihedral (Figure T 8.3), having its origin at the point M
that performs the motion and the axes:
- tangent to the curve, positively oriented in the direction of increase of arc s, of

versor τ ;
- main normal, i.e. normal in the osculating plane of the curve, positively oriented

toward the center of curvature, of versor ν ;

- binormal, i.e. the normal to the osculating plane, of versor β considered so that the
→ → → → → →
versors τ , ν and β taken in this order form a rectangular trihedral ( β = τ × ν ).

binormala

β S 
S  tan genta aτ 
τ M ν
  ( C)
M0 ν ( C) M0 aν 
 a
r

O normala
principala

Figure T 8.3 Figure T 8.4

The motion of the point is given by the hourly law of motion s = s(t) and the
position about the fixed point O is expressed depending on the element of arc s, that is
→ →
r = r (s (t )) (8.4)
In order to determine the velocity and acceleration components on the axes of the
Frenet’s frame, we evoke two formulas from the differential geometry (Frenet's formulas):
→ →
d r → dτ 1→
=τ , = ν (8.5)
ds ds ρ
where ρ is the radius of curvature at point M.
The velocity of the point M is given by the relation:

121
Theoretical mechanics – theory and applications

→ →

d r d r d s •→
v = = ⋅ =sτ (8.6)
dt ds dt
and its projections on the axes of reference under consideration and its magnitude are:
• → •
vτ = s , vν = 0 , v β = 0 , v =s (8.7)

It results that the velocity is tangent to the trajectory, has the sense of the motion
and the scalar equal to the derivative of the arc s with respect to time.

Taking into account the second formula of Frenet and the relation of definition
(8.3), the acceleration vector equals to:

→ → •
→ d v d  • →  • • → • d τ • • → s 2 →
a = = sτ = s τ +s = s τ+ υ (8.8)
dt d t  
 dt ρ

Projections and magnitude of the acceleration vector are:

2
• 2 • 2 
•• s2 →  ••  s  • 4
aτ = s , aυ = , aβ=0 , a =  s  + 2 v
   = v + (8.9)
ρ    ρ  ρ2
 

Acceleration has two components (Figure 8.4), both in the osculating plane
→ → →
(determined by vectors τ and ν ), namely the tangential acceleration a τ (in the direction

of the tangent to the curve (C)) and normal acceleration a ν (in the direction of the main
normal, always oriented towards the inside of the curve). Acceleration vector is therefore
also facing concavity (inside) of the curve.

Note: From the expression (8.9) it results that the only motion in which the acceleration is
zero is the rectilinear and uniform motion (along a straight line at a velocity constant in
magnitude:
→ → • v2
a = 0 ⇒ aτ = aυ = 0 ⇒ v = = 0 ⇒ v = constant, ρ = ∞ (line).
ρ

8.2.2. Cartesian coordinates system



In a Cartesian coordinate system Oxyz the position vector r (t ) is defined by the
coordinates x, y and z of the material point (Figure 8.5). Knowing point motion means
knowing its coordinates as a function of time:
x = x(t) , y = y(t) , z = z(t) (8.10)

122
Theoretical mechanics – theory and applications

Expressions (8.10) are called parametric equations of the trajectory and they allow
obtaining it by eliminating the time variable t. The position vector has the expression:
→ → → →
r (t ) = x(t ) i + y (t ) j + z (t ) k (8.11)
→ → →
where i , j , k are the constant versors of the axes of coordinates.
Analytical expressions, components on the axes and the magnitudes of the velocity
and acceleration vectors in the Cartesian coordinates are:

→ → •→ •→ •→ • • • → • • •
v = r = x i + y j + z k , v x = x, v y = y , v z = z , v = x 2 + y 2 + z 2

••
→ → •• → •• → •• → •• •• •• → •• •• ••
a = r = x i + y j + z k , ax = x , a y = y , az = z , a = x 2 + y 2 + z 2 (8.12)

M
z ( C)

r

y
y
x

Figure T 8.5

y
M

r
  
n ρ
j
θ


O i x

Figure T 8.6

123
Theoretical mechanics – theory and applications

8.2.3. Polar coordinates system

If the trajectory is plane, the point position can be specified by polar coordinates,
 ∧ 
namely by the polar radius r = OM and polar angle θ = mas Ox, OM  measured counter-
 
 
clockwise from Ox (Figure T 8.6). The parametric equations of the trajectory have the
form:

r = r (t ) , θ = θ (t ) (8.13)

To express the components of velocity and acceleration vectors, in addition to


Oxy, a plane movable system of axes having the directions of the polar radius, of
→ →
versor ρ , and perpendicular to the polar radius, of versor n , directed in the increasing
sense of the polar angle θ is considered. Relative to the chosen system, the position vector
is expressed as:
→ → →
r = OM = r ⋅ ρ (8.14)
→ → → →
To find the vectors v and a , the derivatives of the vectors ρ and n are
→ → →
required. Therefore, the versors ρ and n are expressed depending on the versors i and

j as well, and the relations obtained are differentiated with respect to time:
→ → → → → →
ρ = cos θ ⋅ i + sin θ ⋅ j , n = − sin θ ⋅ i + cos θ ⋅ j
• •
→  → → • • → →  → → • • →
ρ =  − sin θ ⋅ i + cos θ ⋅ j  ⋅ θ = θ ⋅ n , n =  − cos θ ⋅ i − sin θ ⋅ j  ⋅ θ = − θ ⋅ ρ (8.15)
   
   
According to the definition, we can write:
• •
→ → •→ → •→ •→
v = r = r ρ + r ρ = r ρ + rθ n
•• • • •
→ → → •• → • → • • → •→
a = r = v = r ρ + r ρ + rθ n + rθ n (8.16)

Taking into account (8.15) we obtain the following analytical expressions of the
speed and acceleration vectors in polar coordinates, as well as their projections on the axes
and their magnitudes.
→ •→ •→ • • → • •
v = r ρ + rθ n , vr = r , vn = r θ , v = r 2+ r 2 θ 2

124
Theoretical mechanics – theory and applications

→  •• •  →  •• • •  →
a =  r − rθ 2  ρ + r θ + 2 rθ  n (8.17)
   
   
2 2
•• • •• •• →  •• •   •• • • 
ar = r − r θ 2 , an = r θ + 2 r θ , a =  r − rθ 2  + r θ + 2 rθ 
   
   

Note: The expressions of the velocity and acceleration vectors were obtained taking into
account the way of expression of the position vector in various reference systems and
• •
→ → → →
using the definition relations v = r and a = v . The intrinsic, Cartesian and polar
coordinates cases were exposed, but other cases may be studied such as that of cylindrical,
spherical, generalized coordinates etc. In all cases the procedure is the same, starting from
the expression of the position vector in that system of coordinates [2].

8.3. Particular motions of the material point:


Rectilinear motion and circular motion

8.3.1. Rectilinear motion

Rectilinear motion is the motion in which the trajectory is a straight line (or part
of a line). To study this motion we can use a Cartesian system with the Ox axis
superimposed on the line of motion (Figure T 8.7). The motion is known if the distance
→ → → • →
OM = x (t) is known at any instant of time. Since r = x ⋅ i it follows that v = x⋅ i ,
→ •• →
a = x ⋅ i i.e. velocity and acceleration vectors are directed on Ox. The motion is
→ →
accelerated (velocity magnitude increases) if the vectors v and a have the same sense,
and decelerated (velocity magnitude decreases) otherwise.

 
ν aτ
M
 •
a  S
x aν
 
ν
•M
a θ
•O
M0

x O

 ω
ε

Figure T 8.7 Figure T 8.8

We shall study below two particular cases of rectilinear motion.

125
Theoretical mechanics – theory and applications

8.3.1.1. Uniform rectilinear motion

Motion of a point on a straight line at a constant velocity is called rectilinear and


uniform.
Denoting by v 0 the velocity scalar, we can write:
• •
v = v 0 = x = constant or x = v 0 = constant (8.18)
Integrating equation (8.18) we obtain x = v 0 t + c 1 where the integration constant
is determined from the initial conditions of motion:
t =0 : x = x0 , v =v0 (8.19)
Thus c 1= x 0 , so that the laws of the uniform and rectilinear motion are:
x = x 0 + v 0 t , v = v 0 = constant , a = 0 (8.20)

8.3.1.2. Uniformly variable rectilinear motion

Motion of a point on a line with constant acceleration is called uniformly variable


rectilinear motion.
Denoting by a 0 the acceleration scalar, we can write:
••
x = a 0 = Constant (8.21)
Integrating twice the equation (8.21) and taking into account the initial conditions
(8.19) we obtain the following law of uniformly variable rectilinear motion:

a0 t2
x = x 0+ v 0 t + , v = v 0 + a 0 t , a = a 0 = constant (8.22)
2
In addition, eliminating the time t between the first two relations (8.22), we obtain
Galilei's formula:
v 2 = v 02 + 2 a 0 (x − x 0 ) (8.23)

8.3.2. Circular motion

Motion of a material point on a circle is called circular motion.


To study the circular motion, Frenet’s frame is considered (Figure T 8.8), whose
axes are:

- Tangent to the circle at M, of versor τ , positively oriented in the direction of
increase of the arc M 0 M = s ;
- Main normal, which is right the normal at M in the plane of the circle (in the

radius direction), of versor υ , positively oriented towards the center of the circle;
- Binormal, i.e. the perpendicular to the circle plane at M.

126
Theoretical mechanics – theory and applications

In the circular motion the law of motion is given by one of the functions:
s = s (t ) or θ = θ (t ) (8.24)
 ∧ 
where s = R ⋅ θ , R being the radius of the circle and θ = mas OM 0 , OM  .
 
 
The velocity scalar is:
• •
v = s = Rθ (8.25)
and the acceleration components are:
• •
•• •• s2
R 2θ 2 •
aτ = s = Rθ , aν = = = Rθ 2 (8.26)
ρ R
Denoting
• dθ ∆θ • dω ∆ω
ω =θ = = lim , ε =ω = = lim (8.27)
d t ∆ t →0 ∆ t d t ∆ t →0 ∆ t

relations (8.25) and (8.27) become :


v = Rω , aτ = Rε , aυ = R ω 2 , a = R ε 2+ ω 4 (8.28)

Quantity ω characterizes the variation of the angle θ per unit of time. It's called
angular velocity and is measured in rad / s. Quantity ε characterizes the variation of the
angular velocity ω per unit of time. It's called angular acceleration and is measured in
rad / s 2 .

8.3.2.1. Uniform circular motion

If the velocity scalar (speed) remains constant (hence the one of the angular
velocity as well) the motion is called uniform circular motion.
Considering the initial condition:
t = 0 : θ =θ 0 (8.29)
we obtain the uniform circular motion laws:

θ = θ 0 + ω 0 t , ω = ω 0 = constant , ε = 0
s = s0 + v 0 t , v = v 0 = constant , aτ = 0 , aυ = R ω 02 (8.30)

where s 0 = R θ 0 , v 0 = R ω 0 .

8.3.2.2. Uniformly variable circular motion

If the scalar of the tangential acceleration remains constant (hence the one of the
angular acceleration as well) the motion is called uniformly variable circular motion.

Considering the initial conditions:


127
Theoretical mechanics – theory and applications

t = 0 : θ =θ 0 , ω =ω 0 (8.31)
we find the following laws of the uniformly variable circular motion:
1
θ = θ 0 + ω 0 t + ε 0 t 2 , ω = ω 0 + ε 0 t , ε = ε 0 = constant (8.32)
2
1 (v + a 0 t )2
s = s0 + v 0 t + a 0 t 2 , v = v 0 + a 0 t , a τ = a 0 = constant, a υ = 0
2 R
where s 0 = R θ 0 , v 0 = R ω 0, a 0 = R ε 0 .

Note: In practice, usually known is the speed (given in revolutions per minute). Between
speed (rev / min) and angular velocity (rad / s) there is the relation:
πn
ω= (8.33)
30

8.4. Solved problems

R 8.1) Given the parametric equations of motion of a material point in Cartesian


coordinates:
π  π 
x = 2 cos t  , y = 5 sin  t  + 3 ,t 0 = 2
4  4 
Required:
a) Determine the trajectory of the point;
b) Determine the velocity and acceleration components of the point at an arbitrary
instant of time and their magnitudes;
c) Determine the radius of curvature of the trajectory and acceleration components
in the intrinsic coordinates at the indicated time instant t 0 .
Solution: a) Eliminating the time between the two parametric equations we obtain
x 2 ( y − 3) 2
the trajectory equation + = 1 , which is the equation of an ellipse with axes
4 25
parallel to Ox and Oy, the center at A (0, 3) and half axes 2 and 5 (Figure R 8.1). The
vehicle departs from B (2, 3) and travels along the ellipse in counterclockwise direction.
b) The speed and acceleration components in the Cartesian coordinate system are:
. π π 5π
. π
v x= x = − sin t , v y= y = cos t
4 2 4 4
.. π π 2 .. 5π 2
π
a x= x = − cos t , a y= y = − sin t
8 4 16 4
and their magnitudes :

π π π 2

v= 4 + 21 cos 2 t , a= 4 + 21sin t
4 4 16 4

128
Theoretical mechanics – theory and applications

π π
.21π 2 sin t cos t
4 4
c) a τ = v = −
16 π
4 + 21 cos 2 t
4
( )
For t 0 = 2 s we get a τ = 0 m / s 2 . At the same instant of time we also have
π 5π 2
v=
2
( m / s) , a =
16
( )
m / s 2 . In determining the radius of curvature we can notice
that:
v2 v2 v2
aν= ⇒ ρ= =
ρ aν a 2 − a τ2
4
So that for t = 2 s it results ρ = m . Finally, the normal component of the acceleration is:
5

5π 2
aν =
v2
ρ
=
16
(
m/s 2 )
y

r
A(0,3) B(2,3)
θ
O
O x

Figure R 8.1 Figure R 8.2

R 8.2) A material point is moving in a plane, the equations of motion with respect to the
polar coordinate system being r = 6 t , θ = 3 t . Determine the trajectory as well as velocity
and acceleration of the point at the instant of time t 1= 2 s .

Solution: By eliminating time t between the parametric equations we obtain the


trajectory equation in polar coordinates, r = 2 θ , representing the spiral of Archimedes
(Figure R 8.2). Velocity and acceleration components are:
• •
v r = r = 6 , v θ = r θ = 18 t (m/s)
and
(m / s 2 )
•• • •• • •
a r = r − r θ 2 = −54 t , a θ = r θ + 2 r θ = 36
respectively, and their magnitudes:

129
Theoretical mechanics – theory and applications

(m / s 2 ) .
→ →
v = 6 1+ 9 t 2 ( m / s) , a = 18 4 + 9 t 2

For t 1= 2 s we find the values:

v r = 6 m / s , v θ = 36 m / s , a r = −108 m / s 2 , a θ = 36 m / s 2 ,
→ →
v = 6 37 m / s , a = 36 10 m / s 2 .

R 8.3) A vehicle starting from rest moves on a straight line and within 60 s reaches a speed
of 12 m / s in a uniformly accelerated motion. Further, the vehicle travels 720 m in a
uniform motion. Finally, the vehicle is uniformly decelerated and stops after a distance of
180 m. Study the motion and draw the diagrams of motion x = x (t), v = v (t) and a = a (t).

Solution: Step I: Uniformly accelerated rectilinear motion


x( t ) = 0,12 t 2 ; v( t ) = 0,24 t ; a ( t ) = 0,24 = constant
Motion lasts t 1 = 50 s and the distance x 1 = 300 m is travelled.

Step II: Uniform rectilinear motion


x( t ) = 300 + 12 t ; v( t ) = 12 = constant; a ( t ) = 0
Motion lasts t 2 = 60 s and the space x 2 = 720 m is travelled.

Step III: Uniformly decelerated rectilinear motion


x( t ) = 1020 + 12 t − 0,2 t 2
; v ( t ) = 12 − 0,4 t ; a ( t ) = −0,4 = constant
Motion lasts t 3 = 30 s and the space travelled is x 3 = 180 m.
Motion diagrams are shown in Figure R 8.3.
a

x
V
1200

1020
0,24

12 50 110 140 t
300
0,40

O O
50 110 140 t 50 110 140 t

Figure R 8.3

130
Theoretical mechanics – theory and applications

R 8.4) From a point on a circle of radius R = 10 m, two vehicles are leaving


simultaneously in opposite directions, with the following motions:
- First moves in the trigonometric sense (the direction of increase of arc s) in accordance
with the hourly law s 1= −3 t 2 + 30 t ( m) ;

- The second moving clockwise in a uniform circular motion at the speed v 2 = 4 m / s .
Determine the instant of time when the two vehicles meet for the first and the second time.

Solution: Let A be the starting point of the two vehicles (Figure R 8.4). The first vehicle
moves initially counter-clockwise but does not keep this direction for too long. Indeed, the

speed at a given instant of time is v 1= s 1= −6t + 30 . It cancels for t = 5 s and then becomes
negative (vehicle moves clockwise). The constant tangential acceleration of the first
•• → →
vehicle motion is a 1τ = s 1 = −6 . Consequently, for t ∈ ( 0,5) , the vectors a 1τ and v 1 are

of opposite directions (uniformly decelerated motion) and for t ∈ (5, ∞) vectors a 1τ and

v 1 have the same sense (uniformly accelerated motion). The normal component of the
v 12 ( − 6t + 30) 2
first vehicle acceleration is a 1ν = = .
R 10

The law of motion of the second vehicle is s 2 = − v 2 = −4 t . The tangential

•• v 22
acceleration is zero, a 2τ = s 2 = 0 and the normal one equals to a 2 ν = = 1,6 m / s 2 .
R
Length of the circumference of the circle is L = 2 π R = 62,8 m . The first vehicle travels in t
= 5 s a space s 1= −3 ⋅ 5 2 + 30 ⋅ 5 = 75 m (larger than the circumference) so that the first
encounter will take place when s 1 + s 2 = 2π R , i.e. − 3 t 2 + 30 t + 4 t = 62,8 . The
7
physically acceptable solution of this equation is t ' = = 2,33 s . The first encounter takes
3
place at point B (the first vehicle travelled 53.44 m and the second one 9.36 m).

In the first 5 seconds, the second vehicle travels 20 m (clockwise) and reaches C "
and the first vehicle reaches C '(arc AC' has 75 - 62.8 = 13, 2 m). From this moment both
vehicles move clockwise and for the next encounter the condition s 1= s 2 arises, i.e.
64
− 3 t 2 + 30t = −4t ⇔ t ' ' = = 21,33 s .At this instant of time the vehicles are at point D
3
(the second vehicle travelled s 2 = 4 t ' ' = 85,33 m )

131
Theoretical mechanics – theory and applications

S1 S2

••
C′• •
B
(prima întâlnire)

R O
•C′′
•D
(a doua întâlnire)
Figure R 8.4

y B

L N

O′ M

A ⋅ d

O x

ω

Figure R 8.5

R 8.5) A right angle shaped bent bar OAB (Figure R 8.5), rotates in its plane with constant
angular velocity ω around the fixed point O. In the same plane there is also the fixed right
line LN so that OO '= d. Knowing that OA = r, determine the velocity and acceleration of
point M along the straight line LN.

Solution: Abscissa of the point M is:


d cos ϕ − r
x = O ' M = AM '−OA' = MM ' ctg ϕ − OA' = (d − r cos ϕ ) ctg ϕ − r sin ϕ = ,
sin ϕ


.
where ϕ = −ω decreases over time).

v=x=
. d − r cos ϕ
⋅ω , a = x = v =
.. . (
2 d cos ϕ − r 1 + cos 2 ϕ )⋅ω 2
sin ϕ
2
sin ϕ
3

132
Theoretical mechanics – theory and applications

8.5. Problems proposed

8.5.1. Conventional tests

TC 8.1) Given the parametric equations of motion of a material point in Cartesian


coordinates:
3
x = 2 − 3t − 6 t 2 , y = 3 − t − 3t 2 , t 0 = 1
2
Required:
a) Determine the trajectory of the point;
b) Determine the components of velocity and acceleration of the point at an
arbitrary instant of time and their magnitudes;
c) Determine the radius of curvature of the trajectory and acceleration components
in the intrinsic coordinates at the indicated instant of time t 0

1 x − x2 −1
TC 8.2) A vehicle runs on the curve of equation y = ln , x > 1, finding at the
2 x + x2 −1
initial instant at point A (1, 0) and having zero velocity. From this moment it begins to
perform a uniformly accelerated motion towards the point B, where it arrives at the speed
v B = 15 m / s after travelling the space s AB = 4 3 m . Required:

a) Vehicle acceleration at point B;


b) Time t 1 it takes to travel the space s AB .

TC 8.3) The parametric equations of motion of a material point in polar coordinates are:
π
ρ = 2R , θ = (1 − sin ω t)
2
where R and ω are constantly positive. Required:
a) Point trajectory;
π
b) Position, speed and acceleration at instant t 0 = .

TC 8.4) A point moves on a circle of radius R = 10 cm as per law s = 5 t 2 − 3t + 2 , where s
is given in centimeters and time t in seconds. Determine the position of the point,
components of the acceleration vector and angle between velocity and acceleration vectors
at time t 1 when velocity magnitude is v 1 = 7 cm / s.

8.5.2. Multiple-choice tests

TG 8.1) A particle P moves with constant velocity V along the curve of equation y = ln x
(m). For what value of abscissa x> 0 the particle acceleration is maximum?
1 1 1
a) x = m ; b) x = m ; c) x = 1 m ; d) x = m .
2 3 2

133
Theoretical mechanics – theory and applications

TG 8.2) A vehicle starting from rest moves along a straight line in a uniformly accelerated
motion. Knowing that after t 1= 10 s it reaches the speed v 1= 5 m / s , what was the space
travelled after t 0 = 1 s ?
a) 0,5 m ; b) 0,25 m ; c) 1 m ; d) 0,75 m.

TG 8.3) Distance AB = 3 m is traveled by an automobile as follows:


- First third, at constant speed V 1 ;
- The second third, at a uniformly accelerated motion, velocity is changing from V 1 to V 2 ;
- Third, at constant speed V 2 .
Determine the average speed V m of the car.
V 1+ 2 V 2 3V 1 V 2 (V 1+ V ) V 2− V 1
a) V m = ; c) V m =
2
; b) V m = ;
1 + V 2+
3 2 2 3
V 4V 1 V 2
3V 1 V 2 (V 2 − V 1 )
d) V m = .
V 12 + V 22 − 2 V 1 V 2

TG 8.4) The minute pointer of a watch is 1.5 times longer than the hour pointer. Calculate
the ratio of the linear velocity of the minute pointer tip and the linear velocity of the hour
pointer tip.
a) 21 ; b) 12 ; c) 15 ; d) 18.

8.6. Indications and answers

TC 8.1) a) Line x – 2 y + 4 = 0. At t = 0 the vehicle finds at A (2 , 3 ) travels to the left.


− 3 − 12 t
, v = (3 + 12 t )
5
b) v x = −3 − 12 t , v y = , a x = −12 , a y = −6 , a = 6 5
2 2
c) For t 0 = 1 the values are :
15 5
, a = 6 5 , a ν = 0 , aτ = 6 5 , ρ = ∞
v=
2
 3
The vehicle finds at point B − 7 ,−  .
 2
v 2B
( )
→ → → 75 3
TC 8.2) a) a B = a τB + a ν
B , where a τB = = m / s 2 and
2 s AB 8

[1 + ( y ' ( x B )) 2 ] 2
3

a νB =
v 2B
=
225
( 2
) , because ρ B = = x 2B = 1 + s 2AB .
y ' ' (x B )
m/s
ρB 49
vB 8 3
b) t 1= τ = ( s) .
a B
15

134
Theoretical mechanics – theory and applications

TC 8.3) a) Since sin ω t ∈ [− 1,1] , ∀ t > 0 , it results that θ ∈ [0 , π ] , ∀ t > 0 . But ρ = 2 R =


constant, so that the point will move on the semicircle of radius 2R situated in the half
plane y ≥ 0 . The motion is an oscillatory motion on this semicircle.
•. •
b) v ρ = ρ = 0 , v θ = ρ θ = −π R ω cos ω t
•• • π 2
Rω 2 •• • ⋅
a ρ = ρ − ρθ 2
=− cos 2 ω t , a θ = ρ θ + 2 ρ θ = π R ω 2 sin ω t
2
π
For t 0 = the following kinematic characteristics are obtained:

Position: ρ = 2 R , θ = 0 ;

Velocity: v ρ = 0 , vθ = 0 , v = 0 ;

Acceleration: a ρ = 0 , a θ = π R ω 2
, a = π Rω 2.

. .
v = s = 10 t − 3 , a τ = v = 10
TC 8.4) v 2 (10 t − 3) 2 a ν  10 t − 3  2
a ν= = , tg ϕ = = 
R 10 a τ  10 

For v = v 1 we get the values:


v 1+ 3
= 1 s , s 1= s (t 1 ) = 4 cm , θ 1=
s1
t 1= = 0,4 rad
10 R
v 1= 7 cm / s , a 1τ = 10 cm / s 2 , a 1 ν = 4,9 cm / s 2 , a = a 12τ + a 12ν = , ϕ 1= arctg 0,49

• 2
τ ν V
TG 8.1) Since a P = v = 0 it results that a =aP = . The acceleration is maximum
P
ρ
1
when the radius of curvature is minimum, i.e. if ρ ' ( x) = 0 . Found x = . Correct answer:
2
a).

TG 8.2) The laws of the uniformly accelerated motion are: x = 0,25 t 2 , v = 0,5 t , a = 0,5 .
For t 0 = 1 s , we get x=0, 25 m. Correct answer : b).

AB 1
TG 8.3) V m = , where t 1= (time to travel the first third),
t 1+ t 2 + t 3 V1
1 2
t 3= (time to travel the last third) and t 2 = (time to travel the second third).
V2 V 1+ V 2
Correct answer: b).

135
Theoretical mechanics – theory and applications

V 1 R 1 ω 1 12 R 1 2π 2π R1
TG 8.4) = = = 18 , because ω 1= ,ω 2= , = 1,5. Correct
V 2 R2ω2 R2 1 12 R 2
answer: d).

136
Theoretical mechanics – theory and applications

9. Kinematics of absolute motion of the rigid solid


9.1. General motion of the rigid

9.1.1. Position parameters of the rigid

Motion of a rigid solid is determined when the position vector, velocity and
acceleration of any point of the rigid in relation to a fixed reference O1 x1 y1 z1 is known at
each instant of time.
→ → →
Let a fixed Cartesian reference O1 x1 y1 z1 be of versors i 1 , j 1 , k 1 with respect to
which the motion of a rigid solid (C) is studied and Oxyz a movable Cartesian trihedral,
→ → →
solidary with the rigid, of versors i , j , k (Figure T 9.1). Choosing the point O as origin of
the movable system is arbitrary.
z
( C)
z1
M    y
• r k j

r1  O
r0 i

k1
x

i1 O1 
ji y1

x1

Figure T 9.1

Let an arbitrary point M be having the position vector r 1 with respect to the fixed
→ →
reference and r with respect to the movable one. Denoting by r 0 the position vector of
the origin O of the movable reference, we can write:
→ → →
r1 = r O + r (9.1)
→ → →
where r 1 , r 0 and r are vector functions of time, assumed continuous, uniform and at
least twice differentiable. Since the solid is rigid, the distance OM is constant during the

motion so that the vector OM will shall have the projections x, y, z on the axes of the Oxyz
→ → →
referential. Instead, the versors i , j , k are functions of time, the axes of the reference Oxyz
being able to change their position during rigid motion. Results:
→ → → →
r =x i + y j+zk (9.2)

137
Theoretical mechanics – theory and applications

→ → → → → → →
where i = i (t ), j = j (t ), k = k (t ) . In order to determine the vector r 1 it is needed to
→ → → →
know the vector functions r 0 , i , j and k . Each of these functions requires knowledge of
three scalar functions, namely the projections of the vector functions on the axes of the
reference Oxyz. The number scalar of unknowns is, therefore, 12. But they are not
independent, between the versors of the movable reference axes the relations being:
→ → → → → → → → → → → →
i ⋅ i = j ⋅ j = k ⋅ k =1 , i ⋅ j = j⋅ k = k⋅ i =0 (9.3)

It results that the vector r 1 (t ) is only expressed by means of six scalar functions

of time, three of these functions coming from the vector r 0 that defines the position of
the origin O of the movable reference system relative to the fixed one. The other three
→ → →
functions come from the versors i , j , k that give direction of the movable system with
respect to the fixed system. It is obtained, thus, that a free in space rigid solid has six
degrees of freedom.
These six independent scalar parameters can be chosen as follows:
- Coordinates of the origin of the movable system with respect to the fixed one:
xO = xO (t ) , yO = yO (t ) , zO = zO (t )
- Euler anglesψ = ψ (t ),ϕ = ϕ (t ),θ = θ (t ) , giving orientation of the axes of the
movable system with respect to those of the fixed system (see Subchapter 9.5).

9.1.2. Distribution of velocities

From relations (9.1) and (9.2) we get:


→ → → → →
r 1 (t ) = r O (t ) + x ⋅ i (t ) + y ⋅ j (t ) + z ⋅ k (t ) (9.4)
To determine the absolute velocity of the point M, relation (9.4) is differentiated
with respect to time:
• • • •
→ → → → → → →
v (t ) = v O (t ) + r (t ) = v O (t ) + x ⋅ i (t ) + y ⋅ j (t ) + z ⋅ k (t ) (9.5)

→ →
where v O = r O represent the absolute velocity of the point O.
• • •
→ → →
For understanding the significance of the derivatives i , j , k , we shall
differentiate relations (9.3) with respect to time:
• • • • • • • • •
→ → → → → → → → → → → → → → → → → →
i ⋅ i = j⋅ j = k ⋅ k = 0 , i ⋅ j + i ⋅ j = j⋅ k + j⋅ k = k ⋅ i + k ⋅ i = 0 (9.6)
We denote
• • • • • •
→ → → → not → → → → not → → → → not
j⋅ k = − j⋅ k = ω x , k ⋅ i = − k ⋅ i = ω y , i⋅ j = − i⋅ j =ω z (9.7)

138
Theoretical mechanics – theory and applications


and consider the scalars ω x , ω y , ω z as being the projections of a vector ω on the axes
of the Cartesian reference Oxyz, namely:
→ → → →
ω =ω x i +ω y j +ω z k (9.8)

We shall discover the significance of this vector later. It is known that projection
of a vector on an axis equals to the scalar product between that vector and the versor of the
 → → → 
axis  pr ∆ v = v ⋅ u ∆  , so that:
 
 
• • • • • •
→ → → → → → → → →
pr Ox i = i ⋅ i = 0 , pr Oy i = i ⋅ j = ω z , pr Oz i = i ⋅ k = −ω y (9.9)

→ → →
i = ω z⋅ j − ω y⋅ k
In addition,
→ → →
i j k
→ → → →
ω × i = ω x ω y ω z = ω z⋅ j − ω y⋅ k (9.10)
1 0 0


→ → →
From (9.9) and (9.10) we find that i = ω × i . Proceeding likewise for the
• •
→ →
derivatives j , k we get the relations:
• • •
→ → → → → → → → →
i = ω× i , j = ω× j , k = ω× k (9.11)
known as Poisson formulas or the equations of motion of the movable trihedral .

Replacing Poisson formulas in the relation (9.5) we find:


→ → → → → → → → → →  → → →
v = v O + x ⋅ω× i  + y ⋅ω× j  + z ⋅ω× k  = v O + ω×x ⋅ i + y ⋅ j + z ⋅ k 
       
       
that is
→ → → →
v = v O + ω× r (9.12)

Relation (9.12) is the formula for determining the distribution of velocities in the

general movement of the rigid (Euler equation for velocity distribution), v represents

the velocity of the arbitrary point M of the rigid, v O the velocity of the origin of the

139
Theoretical mechanics – theory and applications

→ →
movable trihedral, ω is the vector defined by the relation (9.8) and r the position vector
of the point M with respect to the movable trihedral. Comparing relations (9.5) and (9.12)
we get that:

→ → →
r = ω× r (9.13)

Projecting relation (9.12) on the axes of the movable trihedral Oxyz we get the
expressions of the components of the velocity of the arbitrary point M of the rigid:

v x = v O x + ω y ⋅ z − ω z ⋅ y , v y = v O y + ω z ⋅ x − ω x ⋅ z , v z = v O z + ω x ⋅ y − ω y ⋅ x (9.14)

9.1.3. Distribution of accelerations

Differentiating relation (9.12) with respect to time we get the absolute acceleration
of the point M:
• • • •
→ → → → → → →
a = v = v O + ω× r + ω× r

→ not →
But v O = a O represents acceleration of the origin O of the movable trihedral,
• •
→ → → → not →
r = ω × r and ω = ε shall be a vector quantity whose significance we will discuss
later. Thus the formula giving the distribution of accelerations in the general motion of the
solid rigid is obtained. (Euler equation for acceleration distribution):
→ → → → → → →
a = a O + ε × r + ω×ω× r  (9.15)
 
 
Projecting relation (9.15) on the axes of the movable trihedral Oxyz we get the
components of the acceleration of the arbitrary point M of the rigid:
( ) ( )
a x = a Ox + ε y ⋅ z − ε z ⋅ y + ω x ⋅ x + ω y ⋅ y + ω z ⋅ z ⋅ ω x − ω 2 ⋅ x

a y = a Oy + (ε z ⋅ x − ε x ⋅ z ) + (ω x ⋅ x + ω y ⋅ y + ω z ⋅ z ) ⋅ ω y − ω 2 ⋅ y (9.16)
a z = a Oz + (ε x ⋅ y − ε y ⋅ x ) + (ω x ⋅ x + ω y ⋅ y + ω z ⋅ z ) ⋅ ω z − ω 2 ⋅ z
→ → → → → → → →
where ω = ω x i + ω y j + ω z k , ε = ε x i + ε y j + ε z k and ω 2 = ω 2x + ω 2y + ω 2z .

9.1.4. Properties of the velocity distribution in the general motion of the rigid solid

General rigid motion is characterized by the following properties of the


distribution of speeds (without proof):

P 1) Vector ω is an invariant with respect to change of the origin O of the

movable trihedral (likewise vector ε );
140
Theoretical mechanics – theory and applications


P 2) The scalar product of the velocity vector v of an arbitrary point M of rigid and the

vector ω is invariant (does not depend on point M)
→ →
v ⋅ ω = constant (9.17)

P 3) The projections of the velocities of the points of a rigid on the direction of vector ω
are equal:
→ →
→ v⋅ω
pr → v = = constant (9.18)
ω →
ω


P 4) Points situated on a line parallel to the vector ω are of equal velocities.
P 5) The projections of any points M and N of a rigid on the segment joining them are
equal:
→ →
pr MN v M = pr MN v N (9.19)
P 6) The extremities of the velocities of the points of a segment of the solid rigid are
collinear points.
9.2. Particular motions of the rigid solid

Particular motions of the rigid are called those motions wherein, either due to the
mode of action of the system of forces or because of some connections, a part of the six
parameters of position of the rigid remain constant during motion. We address below a
number of five specific motions of the rigid.

9.2.1. Translational motion of rigid

A rigid performs a translational motion if any line stiff-connected to the rigid


remains parallel to itself throughout the motion.

Rigid translations can be rectilinear, circular or arbitrary. Examples of bodies


performing translations are: body of a car on a rectilinear road, bicycle drive pedal,
coupling rod of two wheels of equal radii etc.
In the translational motion the trajectories of rigid points are identical curves that
→ →
can be overlaid with a geometric translation of vector r = AB (Figure T 9.2).
As per the definition, the axes of the movable trihedral Oxyz (solidary with the
rigid) remain parallel to themselves during motion. Result:
• • •
→ → → → → →
i , j , k = constant ⇒ i , j , k = 0 ⇒ ω x = ω y = ω z = 0 ⇒

141
Theoretical mechanics – theory and applications


→ → → → →
ω=0 , ε =ω = 0 (9.20)


ν 3A


a 3A A

ν1A

A
•   
a 3B = a 3A
 
ν 3B = ν 3A

a1A B
  
r ν1B = ν1A 


a 2A
B
 
a1B = a1A
• A t = t3

ν 2A
t = t1

  B
a 2B = a 2A
 
ν 2B = ν 2A

t = t2

Figure T 9.2

From Euler’s equations (9.12) and (9.15) we find that:


→ → → →
v = vO , a = aO (9.21)
which means that at a certain instant all the points of the rigid have the same velocity and
the same acceleration, being sufficient to know the motion of a single point of the rigid so
as to know the motion of all points of the rigid (see Figure T 9.2).
A moving rigid solid has three translational degrees of freedom (displacements
along the axes of coordinates).

9.2.2. Rotation of rigid

9.2.2.1. Generals

A rigid performs a rotation motion if two of its points remain fixed in space
throughout the motion.

The line (∆ ) determined by the two fixed points is also fixed and is called axis of
rotation. Examples of bodies with rotational motions are the rotor of an electric machine,
universal of a lathe, a grinder wheel etc. An arbitrary point M of rigid located off the
rotation axis, describes a circular motion on a circle located in a plane perpendicular to the
axis of rotation, with the center on the axis of rotation and of radius equal to the distance
from point to axis (Figure T 9.3).
For the study of the motion of rotation, a fixed trihedral O 1 x 1 y 1 z 1 and a movable
trihedral Oxyz solidary with the rigid are selected so that O ≡ O 1, O z ≡ O 1 z 1= (∆) . The
angle between axes Ox and O 1 x 1 (or Oy and O 1 y 1 ) is denoted by θ and is the scalar

142
Theoretical mechanics – theory and applications

parameter which sets the rigid position at an instant. Consequently, the rigid solid has only
one degree of freedom in the rotation motion. Rigid motion equation is:

θ = θ (t ) (9.22)
Since O ∈ (∆ ) , it results that:
→ → → →
vO = 0 , aO = 0 (9.23)
(Δ ) − axa de rotatie
z = z1

•M

 
k = k1 y

j
O1 ≡ O θ y1
i1 
  j1
ω
i
x1
θ

ε
x

Figure T 9.3
→ →
In order to determine the vectors ω and ε , the versors of the movable reference
axes are expressed depending on those of the fixed reference (see Figure T 9.3):
→ → →
 i = cos θ ⋅ i 1 + sin θ ⋅ j 1
→ → →

 j = − sin θ ⋅ i 1 + cos θ ⋅ j1 (9.24)
→ →
k = k
 1

Differentiating relations (9.24) with respect to the time t we get:
•
→  → →  • • →
 i =  − sin θ ⋅ i 1 + cos θ ⋅ j 1  ⋅ θ = θ ⋅ j
  
•
→ → →  • • →
 
 j =  − cos θ ⋅ i 1 − sin θ ⋅ j 1  ⋅ θ = − θ ⋅ i (9.25)
  
• •
→ → →
k = k = 0
 1


hence:

143
Theoretical mechanics – theory and applications

• • •
→ → •→ → → → → → → •→ → •
ω x = j ⋅ k = −θ i ⋅ k = 0 , ω y = k ⋅ i = 0 , ω z = i ⋅ j = θ j ⋅ j = θ (9.26)

It results that:
→ • → → •• → • ••
ω =θ⋅ k , ε =θ⋅ k , ω =θ , ε=θ (9.27)
→ →
Vectors ω and ε have the direction of the axis of rotation and their scalars are
obtained by differentiating the function θ (t ) that describes the rigid motion. Since the
→ •
scalar of the vector ω is given through θ , similarly to the circular motion of the material
point, this vector shall be called angular velocity vector and shall characterize variation of

the angle θ . Identically, the vector ε shall be called angular acceleration vector.

9.2.2.2. Study of the distribution of velocities

From (9.12) and (9.23) we obtain for the velocity of an arbitrary point M(x, y, z) of
the rigid under rotary motion, the expression:
→ → →
i j k
→ → → → →
v = ω× r = 0 0 ω = −ω y i + ω x j (9.28)
x y z

The velocity vector projections and magnitude are:


v x = −ω y , v y = ω x , v z = 0 , v = ω x 2 + y 2 = ω d (9.29)
where d is the distance from point M to the axis of rotation.

The properties of the velocity field can be highlighted considering the


points A 1(x 1,0,0 ) , A 2 (x 2 ,0,0 ) , A 3 (x 3,0,0 ) as arbitrarily located on a line perpendicular to
the axis of rotation (Figure T 9.4). The velocities of these points being:
→ → → → → →
v A1 = ω x1 j , v A 2 = ω x 2 j , v A 3 = ω x 3 j
we can deduce the following properties:
i) The only points of zero velocity are belonging to the axis of rotation;
ii) Velocities of different points of the rigid are contained in planes perpendicular
to the axis of rotation;
iii) The points belonging to a line parallel to the axis of rotation have identical
velocities;
iv) The velocities of points lying on a line perpendicular to the axis of rotation are
perpendicular to this line and their magnitudes are directly proportional to the
distance from the point to the axis of rotation.

144
Theoretical mechanics – theory and applications


z  a A′′
v A′′ z 2
2
A′2′
A′2′

a A′
  2
y A′2 v′A
′ ω
2
 y
ω  A′2
   ε
vA vA vA   
1 2 3 aA aA aA
O 1 2 3
O
A1 ϕ
A2 A1
A3 ax ay
x A2
x
A3

Figure T 9.4 Figure T 9.5

9.2.2.3. Study of the distribution of accelerations

From (9.15) and (9.23) we obtain for the acceleration of an arbitrary point M(x, y,
z) of the rigid, the following analytical expression:
→ → → → → →
i j k ij k
→ → → → → →
a = ε × r + ω×ω× r  = 0 0 ε + 00 ω =
 
  x y z −ω y ω x 0

( )→ (
= − yε − x ω 2 i + x ε − y ω 2 j )→ (9.30)

The acceleration vector projections and magnitude are:


a x = − y ε − x ω 2 , a y = xε − yω 2 , a z = 0 , a = d ε 2 + ω 4 (9.31)

The properties of the acceleration field (Figure T 9.5) are also highlighted by the
help of the points A 1(x 1,0,0 ) , A 2 (x 2 ,0,0 ) , A 3 (x 3,0,0 ) for which we have :

→ → → → → → → → →
a A 1 = − x 1ω 2 i + x 1 ε j , a A 2 = − x 2ω 2 i + x 2 ε j , a A 3 = − x 3ω 2 i + x 3 ε j
→ → →
a A1 = x1 ε 2+ ω 4 , a A 2 = x 2 ε 2+ ω 4 , a A 3 = x 3 ε 2+ ω 4

ay ε
tg ϕ = =
= constant (9.32)
ω2 ax
Acceleration distribution properties are identical to those for the distribution of
velocities with the only difference that the accelerations are inclined at the same angle ϕ
with a perpendicular to the axis of rotation (given by the relation (9.32).

145
Theoretical mechanics – theory and applications

9.2.3. Helical motion of the rigid

9.2.3.1. Generals

A rigid is always in a helical motion if in its motion two of its points remain on a
fixed line in space, called helical axis.
Examples of a rigid performing a helical motion are: a drill during the drilling
operation with manual feed, a bullet in the barrel of a rifled gun, a screw etc. Two
trihedrals are chosen for the study of the helical motion, namely the fixed trihedral
O 1 x 1 y 1 z 1 and the movable trihedral Oxyz, stiff-connected to the rigid, so that
O 1 z 1≡ O z ≡ ∆ (helical axis). The origin O of the movable trihedral moves on the fixed
axis O 1 z 1 (Figure T9.6). At the initial instant (t =0) it can be considered that O ≡ O1 . In the
helical motion the rigid has only two degrees of freedom since its constraints allow a
translation along the axis (∆ ) and a rotation around the same axis. The motion of the rigid
is a combination of two simultaneous independent motions, the motion parameters being
the distance z O = z O (t ) of the point O to the fixed trihedral and the angle θ = θ (t )
between the axes O 1 x 1 and Ox (or O 1 y 1 and Oy).

z = z1 z = z1


  ν
ν0
M k
y   
ν = ω× r

r  M
O j O1y1  y
 ν0
 r
i z0
O
x1
O1x1 θ 
O1 ω
y1
x
x O1

y1 x1

Figure T 9.6 Figure T 9.7

Since the origin of the movable system has a rectilinear motion on O 1 z 1 , we can
write:
→ • → → → •• → →
v O = zO k =v O k , a O= z O k = a O k (9.33)

Oxy plane always remains parallel in motion to the plane O 1 x 1 y 1 , so that the
→ → →
versors i , j and k of the movable trihedral have the same expressions and derivatives as
with the movement of rotation.
Therefore:
→ •→ → → •• → →
ω =θ k =ω k , ε = θ k =ε k (9.34)

146
Theoretical mechanics – theory and applications

9.2.3.2. Study of the velocity distribution

Velocity distribution in the helical motion is achieved by the general formula


(9.12) which in this case becomes:
→ → →
i j k
→ → → → → → → →
v = v O + ω× r = vO k + 0 0 ω = −ω y i + ω x j + v O k (9.35)
x y z

The velocity vector projections on the movable reference and its magnitude are:
v x = −ω y , v y=ω x , v z= v O ,v = v O 2 + ω 2 x 2+ y 2 ( (9.36) )
It can be remarked (see also subchapter 9.2.2) that the velocity distribution is
obtained by superposition of two velocity fields: first one corresponding to a rotation

around the axis Oz at an angular velocity ω and the second one specific to a translation

along the axis Oz at velocity v O (Figure T 9.7). The properties of the velocity distribution
are:
i) Generally, there is no zero velocity point. The points of minimum velocity
(equal to v O ) belong to the helical axis;
ii) Rigid points located on a parallel line to the helical axis have the same speed;
iii) Rigid points belonging to a perpendicular line to the helical axis have velocity
magnitudes directly proportional to the distance from the point to the axis (∆ ) .

9.2.3.3. Study of acceleration distribution

Using the general equations (9.15) and the relations (9.33-9.34) we obtain :

→ → → → → →
i j k i j k
→ → → → →  → → →
a = a O + ε × r + ω×ω× r  = aO k + 0 0 ε + 0 0 ω =
 
  x y z −ω y ωx 0

( )→ ( )→
= − y ε − xω 2 i + xε − yω 2 j + a O k

(9.37)

Components on the axes and the magnitude of the acceleration of the arbitrary
point M(x, y, z) are:
a x = − y ε − xω 2 , a y = x ε − y ω 2 , a z = a O , a = a O ( )(
2 + ω 4+ ε 2 x 2+ y 2 (9.38))
Acceleration distribution results as well as a superimposition of vector fields
(acceleration), the first one corresponding to a rotation around the axis Oz at the angular

acceleration ε and the second one specific to a translation along Oz, at the acceleration

147
Theoretical mechanics – theory and applications


a O . The properties of the acceleration distribution are the same as with the velocity
distribution.

9.2.3.4. Particular case: screw displacement

A particular case of the helical motion, very common in engineering, is the screw
displacement. The particularity is that with a full rotation, the rigid advances with a
constant pitch p. There is, thus, a dependency of the type:
z O (t ) = k ⋅ θ (t ) (9.39)
between the two parameters characterizing the helical motion, the rigid having only one
degree of freedom. Imposing the condition:
θ = 2π : z O= p (9.40)
p
we obtain k = . Following successive derivations of (9.39) the following relations

linking the cinematic elements result:
p p
v O= ω , a O= ε (9.41)
2π 2π

9.2.4. Plane-parallel motion of the rigid

9.2.4.1. Generals

A rigid has a plane - parallel motion if three non-collinear points thereof remain
in a fixed plane in space throughout the motion. The plane determined by the three points
and tied to the rigid remains in the fixed plane, too.

Examples of items performing a plane-parallel motion are: the connecting rod in


an engine, the wheel of a vehicle moving on a straight road etc. From geometrical
considerations it follows that all points of the rigid have trajectories contained in planes
parallel to the fixed one. Any point of the rigid remains at a constant distance from the
fixed plan and has the same motion as its projection on the fixed plan. In order to know the
motion of all of the points of the rigid it will be sufficient to determine the motion of the
fixed points of the rigid in the fixed plane.

For the study of plane-parallel motion let us consider the fixed trihedral O 1 x 1 y 1 z 1
for which the plane O 1 x 1 y 1 is even the fixed plane and the movable trihedral Oxyz, stiff-
connected to the rigid, wherein the plane Oxy is invariably connected to the section
determined in the rigid by the fixed plane (Figure T 9.9). The rigid position at a certain
instant is completely determined if the coordinates of the origin O of the movable
reference and the angle θ between axes Ox and O 1 x 1 (or Oy and O 1 y 1 ) are known, that is
by the functions:
x 1O = x1O (t ) , y 1O = y 1O (t ) , θ = θ (t ) (9.42)

148
Theoretical mechanics – theory and applications

It results that the rigid has three degrees of freedom in the plane-parallel motion.
z1
z

k1

j1 M y1
  
i1 O1 r k y
 θ
O j
 O1y1
x1 O1x1 i
plan fix
θ
x

Figure T 9.9

Since point O moves only in the plane Oxy (or O 1 x 1 y 1 ) the velocity v O and

acceleration a O vectors have components in this plan only:
→ • → • → → →
v O = x 1O i 1+ y 1O j 1= v O x i + vO y j
→ •• → •• → → →
a O = x 1O i 1+ y 1O j 1= a O x i + aO y j (9.43)
• • •• ••
Note: v O x ≠ x 1 O , v O y ≠ y 1O , a O x ≠ x 1O , a O y ≠ y 1 O , since in (9.43) projections
materialize on two different trihedrals.
→ → →
The projections of the versors i , j and k on the axes of the fixed
reference O 1 x 1 y 1 z 1 remain the same as with the rotation motion so that:
→ •→ → → •• → →
ω =θ k =ω k , ε = θ k =ε k (9.44)

9.2.4.2. Study of the velocity distribution

Taking into account the general formula (9.12) and the relations (9.43) and (9.44)
the following analytical expression is obtained for the velocity of the arbitrary point
M(x,y,z) :
→ → →
i j k
→ → → → → →
v = v O + ω × r = vO x i + vO y j + 0 0 ω =
x y z

→ →
( )
= v Ox− ω y i + v O y+ ω x j ( ) (9.45)

149
Theoretical mechanics – theory and applications

Point M velocity projections on the axes are:


v x = v O x − ω y , v y = vO y + ω x , v z = 0 (9.46)

Any point of the rigid has the velocity contained in a plane parallel to the plane
Oxy. Velocity distribution in plane-parallel motion of rigid has the following properties:
i) There are, in general, points of zero velocity. Denoting by ξ , η , ζ the
coordinates of such a point and equating to zero the components of the velocity given by
(9.46) we obtain:
v v
ξ = − O y , η = O x , ζ = arbitrary (9.47)
ω ω
Relations (9.47) represent the parametric equations of a line perpendicular to the
plane Oxy that is called instantaneous axis of rotation. Point I (ξ , η ,0 ) given by the
intersection between this line and the plane Oxy is called instantaneous centre of rotation
(ICR) and is the only point in the plane Oxy whose velocity is null at a given instant.
The instantaneous centre of rotation (ICR) is a variable point both with respect to
the fixed trihedral O 1 x 1 y 1 z 1 and to the movable trihedral Oxyz. Continously, another
point of the rigid has zero velocity and becomes ICR. The geometrical locus of the ICR
with respect to fixed trihedral is a curve called base (or fixed centrode) and the geometrical
locus of the ICR with respect the movable trihedral is a curve called roller (or movable
centrode). It can be demonstrated that during the rigid motion the movable centrode rolls
over the fixed centrode, the two curves being tangent at the ICR at that instant. In space,
the instantaneous axis of rotation generates two cylindrical surfaces with respect to both
the fixed and movable trihedrals, called fixed axode and movable axode. The movable
surface rolls without slipping over the fixed one, having in common the instantaneous axis
of rotation at every instant of time.
ii) The distribution of velocities in the plane-parallel motion is identical to that of
rotation motion, as if the solid would rotate about the instantaneous axis of rotation with

the angular velocity ω .
Indeed, considering that at that instant the origin O of the movable trihedral coincides with
→ → → → → →
ICR (O ≡ I ) , the velocity of point M will be v M = v I + ω × IM = ω × IM
→ →
(because v I = 0 ), identical to that of rotation about I.

9.2.4.3. Study of acceleration distribution

The acceleration of the arbitrary point M(x, y, z) results from the general formula
(9.15) and the relations (9.43) and (9.44):
→ → → → → →
i j k ji k
→ → → → →  → → → →
a = a O + ε × r + ω ×  ω × r  = aO x i + aO y j + 0 0 ε + 0 ω =
0
 
  x y z −ω y ω x 0

150
Theoretical mechanics – theory and applications

( )→ (
= a Ox− ε y − ω 2x i + a O y+ ε x − ω 2 y j )→ (9.48)

Point M acceleration projections on the axes of the Oxyz reference are:


a x = aO x − ε y − ω 2 x , a y = aO y + ε x − ω 2 y , a z = 0 (9.49)
Any point of the rigid has the acceleration contained in a plane parallel to the plane
Oxy. Acceleration distribution in plane-parallel motion of rigid has the following
properties:
i) There are, in general, points of zero acceleration. Denoting by ξ ' , η ' , ζ ' the
coordinates of such a point and equaling to zero the components of the acceleration, it
results:
ω 2 a Ox − ε a O y ε aO x + ω 2 a O y
ξ '= , η '= , ζ '= 0 (9.50)
ω 4+ ε 2 ω 4+ ε 2
The relations (9.50) represent the parametric equations of a line perpendicular to
the plane Oxy, at the points of which the accelerations are zero at the given instant. The
intersection of this line with the plane Oxy is the point J (ξ ' , η ' ,0 ) , called the pole of
accelerations. It is the single point of acceleration zero in the fixed plane at a given instant.
ii) The distribution of accelerations in the plane-parallel motion is identical to that in
the rotation motion, as if the rigid would rotate around an axis normal to the fixed plane
passing through the pole of accelerations.
Indeed, if considered at that instant that the origin O of the movable trihedral coincides
with J (O ≡ J ) , then the acceleration of point M will be:
→ → → → → → →  → → → → → 
a M = a J + ε × JM + ω ×  ω × JM  = ε × JM + ω ×  ω × JM 
   
   
→ →
(as a J = 0 ), identical to that of the rotation motion about J.

Note : The ICR, I and the pole of accelerations J are different points at a given instant
→ → → → → → → →

 v I = 0 but v J ≠ 0 and a J = 0 but a I ≠ 0  .
 

9.2.5. Spherical motion of the rigid

9.2.5.1. Generals

A rigid has a spherical motion when one of its points remains confounded with a
fixed point in space throughout the motion.
From geometrical considerations it follows that all points of the rigid have
trajectories contained on spheres with the center at the fixed point and radius equal to the
distance from the corresponding point to the fixed point.
Examples of rigid solids with a spherical motion are the gyroscope, the motion of a
ball joint device on a tripod etc.

151
Theoretical mechanics – theory and applications

For the study of the motion, two references are taken, namely a fixed one
O 1 x 1 y 1 z 1 and a movable one Oxyz stiff-connected to the rigid, so that the origins O 1 and
O are confounded with the fixed point (Figure T 9.10). The position of the rigid is
completely determined at a given instant by Euler anglesψ , ϕ , θ (defined below), so that in
the spherical motion the rigid has three degrees of freedom.
z1

z M

 
θ r ω
 y
ε
y1
O1 = O

x1 ψ
ϕ

Figure T 9.10
Intersection of the planes Oxy and O 1 x 1 y 1 is the line ON, called line of nodes.
Euler angles are independent of each other and are defined as follows:
- precession angle ψ , formed by the line of nodes with the axis O 1 x 1 ;
- eigenrotation angle ϕ , formed by Ox with the line of nodes ;
- nutation angle θ , formed by Oz with O 1 z1 .
Since origin O of the movable reference is fixed it follows that:
→ → →
v O= a O= 0 (9.51)
→ →
Vector ω and vector ε have components onto all axes of the movable trihedral
(see chapter 15), their analytical expressions being of the form:
→ → → → → → → →
ω =ω x i +ω y j +ω z k , ε =ε x i +ε y j +ε z k (9.52)

9.2.5.2. Study of the velocity distribution

Velocity of the arbitrary point M(x, y, z), given by (9.12), (9.51) and (9.52), is:
→ → →
i j k
→ → →
v =ω × r =ωx ωy ωz ⇔
x y z

→ → → →
( )
v = ω y z − ω z y i + (ω z x − ω x z ) j + ω x y − ω y x k (
(9.53) )
Projections of the velocities of the point M on the axes of the movable reference
are:

152
Theoretical mechanics – theory and applications

v x= ω y z − ω z y , v y=ω z x −ω x z , v z=ω x y −ω y x (9.54)

Velocity distribution in the spherical motion has the following properties:


i) There are, in general, points of zero velocity. Equating to zero the velocity
components given by (9.54), we obtain a homogenous linear system with the unknowns x,
y, z, that is simply undetermined compatible and has the solution given by x = λ ω x ,
y = λ ω y , z = λ ω z , λ ∈ R , which can be written as:
x y z
= = (9.55)
ωx
ωy ωz
Zero velocity points are on a line containing the fixed point O(0, 0, 0) and has the

direction of the vector ω (directing parameters ω x , ω y , ω z ). This line is called the
→ →
instantaneous axis of rotation and is a line variable in time, since ω = ω (t). The
geometrical locus of the instantaneous axis of rotation relative to fixed trihedral is a
conical surface with the vertex at O 1 , called herpolodia cone and the geometrical locus of
the instantaneous axis of rotation in relation to the movable trihedral is a conical area with
the vertex at O and named polodia cone. During the motion the herpolodia cone remains
fixed and the polodia cone rolls without slipping over the herpolodia cone.

ii) Velocity distribution is the same as in the case of rotational motion, as if the rigid
would rotate about the instantaneous axis of rotation (see equation (9.53)).

9.2.5.3. Study of the acceleration distribution

Using the general formula (9.15) and the relations (9.51) and (9.52) we obtain the
following projections for the acceleration of the point M(x, y, z):

( ) ( ) (
a x = − ω 2y + ω 2z x + ω y ω z − ε x y + ω xω z + ε y z )
a y = (ω y ω x + ε z ) x − (ω 2x + ω 2z ) y + (ω y ω z − ε x ) z (9.56)

a x = (ω z ω x − ε y ) x + (ω z ω y + ε x ) y − (ω 2x + ω 2y ) z

Providing that acceleration of the point M equals to zero (a x = a y = a z = 0 ) a


homogeneous linear system of equations in the unknowns x, y and z is obtained. In order
for this system to admit nonzero solutions it is required that the determinant of the system
cancels, that is:
(
− ω 2y + ω 2z ) ω y ω z − ε x ω xω z + ε y
( )
2
 → →
∆ = ω yω x + ε z 2 2 
− ω x + ω z ω yω z − ε x = − ω × ε  (9.57)
 
ω zω x − ε y ω zω y + ε x (
− ω 2x + ω 2y )  

→ →
Relation (9.57) is only verified if the vectors ω and ε are collinear or one of

153
Theoretical mechanics – theory and applications

them is null. But since, in general, these vectors are non-zero and have different lines of
action, the only point of zero acceleration is the fixed point O (0, 0, 0).
In conclusion, in the spherical motion the distribution of accelerations is specific
and can not be reduced to that corresponding to another particular motion of the rigid.

9.3. General motion of the rigid

9.3.1. Generals

In section 9.2 we focused on a number of five particular motions of the rigid.


Particular motions are characterized by imposing restrictions of geometric nature on the
rigid, which results in reducing its degrees of freedom.
In this chapter we shall recall some aspects of the distribution of velocities and
accelerations in the general motion of the rigid. Absence of the geometric constraints gives
→ → →
the rigid the maximum number of degrees of freedom, i.e. six. Vectors v O , a O , ω and

ε are ordinary functions of time and have the analytical expressions:
→ → → → → → → →
v O = v O x i + v O y j + v Oz k , a O = a O x i + a O y j + a Oz k
→ → → → → → → →
ω =ω x i +ω y j +ω z k , ε =ε x i +ε y j +ε z k (9.58)

The lines of action of the four vectors have arbitrary directions in space (in the
most general case).

9.3.2. Study of the velocity distribution

Velocity distribution in the general motion of the rigid is given by the Euler
equation:
→ → → →
v = v O + ω× r (9.59)
Velocity vector components on the axes are:
v x = v O x + ω y z − ω z y , v y = v O y + ω z x − ω x z , v z = v O z + ω x y − ω y x (9.60)

We shall demonstrate that the instantaneous velocity distribution in the general


motion is reducible to the one specific to the rigid helical motion. For this it is enough to
→ →
show that there are points where the vectors v and ω are collinear, i.e.:
→ →
v = λ ⋅ω ,λ∈R (9.61)
property analogue to that in the helical motion (see section 9.2.3).
Condition (9.61) is equivalent to the relation:
vx vy v
= = z (9.62)
ωx ωy ωz

154
Theoretical mechanics – theory and applications

or, after replacing the velocity vector components, with :


v Ox+ ω y z − ω z y v Oy+ ω z x −ω x z v Oz+ ω x y − ω y x
= = (9.63)
ωx ωy ωz
Equations (9.63) represent equations of a line parallel to the line of action of the

vector ω . It is called instantaneous axis of the helical motion.
Thus, in the general motion of the rigid the velocity distribution at a certain instant
is obtained as if the rigid performs a helical motion about the instantaneous axis of helical

motion at the angular velocity ω and the translation velocity:
→v O + ω× r  ⋅ ω → →
→ → →
→ →
v⋅ ω   vO⋅ω

v= =  = (9.64)
i → → →
ω ω ω
Velocity of an arbitrary point M of the rigid in general motion has two
→ →
components: one, equal to v i , is parallel to ω and independent of the point, and the
→ →
second one, denoted v p , depends on the point and is perpendicular to ω (see Figure

T9.11). Instantaneous helical motion axis points have the minimum velocity, equal to v i

because the component normal to ω is zero.
The geometrical locus of the instantaneous axis of helical motion with respect to
the fixed reference is a ruled surface called fixed axode and the geometrical locus of the
same axis with reference to the movable reference (stiff-connected to the rigid) is also a
ruled surface called movable axode.
Axa instantanee
a mişcării elicoidale


 ν
νi

 
 ν i = ν min
M νp


ω

Figure T 9.11

The two axodes are tangent to the instantaneous axis of the helical motion and the
movable axode rolls over the fixed axode around the common tangent while there is also a

sliding along it at the velocity v i .

155
Theoretical mechanics – theory and applications

Concluding, the idea emerges that the distribution of velocities is specific to each
individual motions. It can be reduced to an instantaneous velocity distribution specific to
the translational motion, rotation motion (for plane-parallel motion or spherical motion)
or helical motion (for the general motion).

9.3.3. Study of the acceleration distribution

Distribution of accelerations in the general rigid motion is given by Euler's


equation:
→ → → → →  → →
a = a O + ε × r + ω×ω × r  (9.65)
 
 
Components on the axes of the movable reference Oxyz are given by the relations
(9.16). To investigate whether the distribution of accelerations in the general motion is
reducible to the motion corresponding to a particular movement, the existence of zero
acceleration points is searched. Assuming that a x = a y = a z = 0 , from (9.16) we obtain a
non-homogeneous linear system of equations in the unknowns x, y and z. The determinant
of the system (identical to that obtained in spherical motion) is:
2
 → →
∆ = − ω × ε  (9.66)
 
 
→ →
In general, the vectors ω and ε are not collinear, hence ∆ ≠ 0 . Therefore the
distribution of accelerations in the general motion of the rigid is similar to the spherical
motion, i.e. there is a unique zero acceleration point (of coordinates equal to non-
homogeneous linear system solution) called acceleration pole. This point changes its
position with respect to both the fixed reference and the movable reference.
But, if ∆ = 0 , then the following cases are possible:
i) there is no zero acceleration point, this being the case of translation and helical
motions;
ii) there is an infinity of non-zero acceleration points located on a line, this being
the case of the rotation or plane-parallel motions.
In conclusion, the distribution of accelerations is specific to each particular motion in part.

9. 4. Solved problems

R 9.1) Consider the wheels O and O’, coupled by means of the arms O’A’ and O”A” and
the rod A’A”, respectively (Figure R 9.1). Knowing that the locomotive is running at the
constant speed v 0 , find:
a) Velocity of an arbitrary point M of the rod;
b) Trajectory described by the point M.
Solution: a) Throughout the motion the arms O’A’ and O”A” remain parallel, so
that the tetragonal O’A’A”O” is in fact a parallelogram. Therefore, the sides A’A’’ and
O’O” are parallel, too, during the motion and since the line O’O” is fixed in space, the rod
→ →
motion is a translation. It results that v M = v A '. The motion of the wheel with the centre

156
Theoretical mechanics – theory and applications

y 
V0
M
A′ A′′

r r
θ x
O′
O′′

Figure R 9.1
v0
O’ is a plane-parallel motion of angular speed ω = , the instantaneous center of rotation
r
being at point I. The velocity of the point A’ has the characteristics:

⊥ IA '
→ 
v A i  trigonometric sense
 90 + θ
v A ' = ω ⋅ IA ' = 2 r sin = v 0 2 (1 − sin θ )
 2

b) The Cartesian coordinates of the point M with respect to fixed reference O1 x 1 y 1 are:
x=v 0 t + A ' M + r cos θ , y = r sin θ ,
v0
Where θ = t , and represent the parametric equations of a cycloid (translated with the
r

vector A ' M with respect to the cycloid described by the point A’ )

R 9.2) A homogenous flat plate in the shape of a disc of radius R rotates around a fixed
axis ( ∆ ) perpendicular to the plate plane (Figure R 9.2.1). Knowing the speed
→ →
v A = λ AB ( λ > 0) of a point on the periphery of the wheel, the direction of the velocity of

the point B ( v B ⊥OB ) and the distance AB = d, determine:
a) Point where axis ( ∆ ) intersects the plate plane;
b) angular velocity ω of the rotation of the plate;
c) velocity of the point B (magnitude and sense);
d) velocity of the plate center O.

Solution: a) Let I be the point where the rotation axis intersects the plate plane
→ →
(Figure R 9.2.2). Since v A ⊥IA, v B ⊥IB , it follows that I is at the intersection of the line
OB with the perpendicular to AB at A. ∆OCB ≈ ∆IAB ⇒ IB = 2 R (point I belongs to the
disc periphery).

157
Theoretical mechanics – theory and applications

A
⋅ 
A I VA

R
C
O ⋅

× B 
R
O V0

⋅ B


VB

Figure R 9.2.1 Figure R 9.2.2

 
 ⊥ IB
⊥ (OAB ) →  → 1→

  →
b) ω clockwise sense ; c) v B sense given by ω d) v O = v B
  2
v λd 2R
ω = A = v B = IB ⋅ ω = λd
 IA 4R 2− d 2  4R 2 − d 2

R 9.3) A rectangular parallelepiped [OABCDEFG] with the sides OA = 3 m, OB = 4 m,


OC = 5 m rotates about the diagonal OF at the angular speed ω = 2 rad / s (Figure R 9.3).
Determine velocities and accelerations of the points F, A and D.
z

C E

G F


ω
B y
O

A D

Figure R 9.3

Solution: The angular velocity vector has the analytical expression:



→ → OF 3 2 → 4 2 → →
ω = ω ⋅ u OF =ω⋅ = i+ j+ 2 k .
OF 5 5
→ → →
Since point F belongs to the rotation axis it results that v F = a F = 0 . For
velocities and accelerations of points A and D we obtain:

158
Theoretical mechanics – theory and applications

→ → →
i j k
→ → → 3 2 4 2 → 12 2 →
v A = ω × OA = 2 =3 2 j− k
5 5 5
3 0 0

→ → →
i j k
→ → → 3 2 4 2 → →
vD = ω × OD = 2 = −4 2 i + 3 2 j
5 5
3 4 0

→ → →
i j k
→ → → → → → 3 2 4 2 246 → 72 → 18 →
a A = ω × (ω × OA) = ω × v A = 2 = i+ j+ k
5 5 25 25 5
12 2
0 3 2 −
5

→ → →
i j k
→ → → → → → 3 2 4 2 → → →
a D = ω × (ω × OD) = ω × v D = 2 = −6 i − 8 j + 10 k .
5 5
−4 2 3 2 0

Magnitudes of these vectors are:

v A = 5,43 m / s , v D = 7,07 m / s a A = 17,80 m / s 2 , a D = 14,14 m / s 2 .

R 9.4) The radius of a screw is r and the thread inclination angle is α . Knowing that the
screw rotates in the nut at the angular velocity ω , find the speed v of a point on the screw
axis.
Solution: The speed v of an arbitrary point on the periphery of the screw (in helical
ω
motion) is equal to v = n ⋅ p , where n = is the speed of the motion and p = 2 π r tg α is
π
the pitch of the screw. Units used are: [v] = m / s [r] = m, [ ω ] = 1 / s

R 9.5) A bar AB = l is moving in the plane O 1 x 1 y 1 so that the end A slips along the axis

O 1 x 1 at the velocity v A . The bar is tangent at C to the circle of radius r and center
O 1 (Figure R 9.5.1). Requested:

a) Movable and fixed centrodes in the plane-parallel motion of the bar;


b) Speed of the points B and C.

159
Theoretical mechanics – theory and applications

I

ω
y1

y1 y
B
B

 x
C
VB C 

VC
 θ 
r VA
A θ ⋅
VA
O1 x1 O1 A x1

Figure R 9.5.1 Figure R 9.5.2


Solution: a) Since the bar is always tangent to the circle C (O 1, r ) it follows that the
velocity of the tangency point C will be along the bar AB (normal to the radius O 1C ). To
determine the centrodes of the plane-parallel motion described by the bar it is required to
determine the position of the instantaneous center of rotation (ICR) for an arbitrary
position of the bar, given by the angle θ (see Figure R 9.5.2). ICR is located at the
intersection of the perpendiculars to the lines of action of the velocities of the points A and
C, the perpendiculars being drawn at these points.

Determining the fixed centrode

ICR coordinates about the fixed reference system O 1 x 1 y 1 are:


r sin θ
x 1= O 1 A = , y 1= IA = O 1 A ⋅ tg θ = r
cos θ cos 2 θ
Eliminating the parameter θ between these two relations, the fixed centrode
equation is obtained:
r 2
(x 12 + y 12 ) − x 14 = 0 .
Determining the movable centrode

Let the movable reference system Axy be, stiff-connected to the bar. The ICR
coordinates about this coordinates system are:
x = IC = AI sin θ = r tg 2 θ , y = AC = r tg θ .
Equation of the movable centrode, obtained by eliminating the parameter θ
between the latter two relations is y 2 = r x , i.e. equation of a parabola having the axis Ox
as axis of symmetry.
v A v A cos 2 θ
b) Magnitude of the angular velocity is ω = = , the sense of this vector
IA r sin θ

being given by the velocity v A (trigonometric sense). Velocity of point C is a vector

160
Theoretical mechanics – theory and applications

directed along the line AB, from B to A and has the magnitude vC = IC ⋅ ω = v A sin θ and
the velocity of the point B is a vector perpendicular to IB, of magnitude
sin θ
v B = IB ⋅ ω = ω IA 2 + AB 2 − 2 IA ⋅ AB cos θ , where IA = r and AB = l.
cos 2 θ

R 9.6) A disk of radius r rolls without slipping on a horizontal plane (Figure R 9.6.1).

Velocity of the disk center is v O . At a point A on the periphery of the disc the bar AB of
length l is hinged, whose end B moves on a horizontal plane. Calculate the velocity of the
points A and B depending on the angle of rotation θ of the disc.

I2 
1
  VA
O V0 ω2 
O V0
A
l θ r 2
l A θ r
B B ϕ
⋅ 
θ/2

I1 ω1
VB

Figure R 9.6.1 Figure R 9.6.2

Solution: The disk and the bar have plane-parallel motions. The disk rolls without
slipping over the horizontal plane so that the ICR in plane-parallel motion of the body is at
the point I 1 of contact between the disk and the plane (Figure R 9.6.2). It follows that:
 
⊥ (OAB ) ⊥ I 1 A
→  →  →
ω 1 clockwise sense ; v A sense given by ω 1
 v  θ
ω 1= 0 v A = I 1 A ⋅ ω 1= 2 v 0 sin
 r  2
 → 
ICR of the body 2, I 2 , is at the intersection of the line I 1 A ⊥ v A  with the
 
 → 
vertical in B ⊥ v B  so that:
 
 
⊥ (OAB ) ⊥ I B
→  →  2

ω 2  trigonometric sense ; v B  to the right .


 v  I B
ω 2 = A v B = I 2 B ⋅ ω 2 = 2 v A
 I 2A  I 1A

161
Theoretical mechanics – theory and applications

 ∧ 
To determine the distances I 2 A and I 2 B denote mas ABI 1  = ϕ and apply the
 
theorem of sines in triangles ABI 1 and ABI 2 . It is found that:
AB I 1A AB I 2A I 2B
∆ ABI 1 : = ; ∆ ABI 2 : = =
θ sin ϕ  (
θ  sin 90 − ϕ )  θ
.
sin sin90 −  sin180 − ϕ − 
2  2  2
Solving this system we obtain the distances:
1 4θ
I 2A= l 2 − 4 r 2 sin
θ 2
cos
2
2r 
I 2B =
1  sin 2 θ cos θ + 1 sin θ l 2 − 4r 2 sin 4 θ  .
θ l
cos 
2 2 l 2 2
2

R 9.7) Let the system of bodies be in Figure R 9.7.1., for which the distances
O 1 A = AB = BC = BD = DE = EF = l , CD = l 1 and angles α , β are given. Knowing that
the crank O 1 A rotates with angular velocity ω , determine the velocities of points A, B, C,
E and F and angular velocities of the elements of the mechanism.

A C VC

ω2 i2 ⋅
C
1 2 3 
VA
ω α B
O1 ⋅
x x A
6

ω1
α α
O1
B
1

D
γ
β  
ω3 = ω4 
i3 = D VE
E β ⋅
E
5 
VF

7 F β ω5
i5
F

Figure R 9.7.1 Figure R 9.7.2

Solution: The seven bodies of the system have the following motions:
Body 1 - rotation motion;
Body 2 - plane – parallel motion;
Body 3 - plane – parallel motion;
Body 4 - rotation motion;
Body 5 - plane – parallel motion;

162
Theoretical mechanics – theory and applications

Body 6 - translation motion;


Body 7 - translation motion.

Point A, considered as point of the crank O 1 A executes a circular motion on a



circle with the center at O 1 and of radius O 1 A = l , at the velocity v A, normal to O 1 A

having the sense of the angular velocity ω (trigonometric) and the
magnitude v A = O 1 A ⋅ ω = l ω .
The AB bar ICR is at the intersection of the vertical line at B with the line O 1 A

(see Figure R 9.7.2). Angular velocity ω 2 in the plane-parallel motion of the bar has a
vA
clockwise sense and the magnitude ω 2 = = ω as I 2 A = O 1 A = l .
I 2A
Velocity of point B (and velocity of any point of the body 6) is a vector
perpendicular to I 2 B , oriented to the left, and has the
magnitude v B = I 2 B ⋅ ω 2 = 2 l ω sin α .
Velocity of point C, considered as a point of the triangular plate CDE, is a vector
perpendicular to the radius CD so that the ICR I 3 ≡ D of the bar BC will be at the
intersection of the straight line CD with the normal in B on O 1 B , that is the angular

velocity ω 3 of plane-parallel movement of the bar BC, is determined as sense and
magnitude by the velocity of point B. We obtain:
v
ω 3 = B = 2 ω sin α , v C = CD ⋅ ω 3 = 2 l 1 ω sin α
BD
CDE triangular plate rotation will be characterized by angular velocity vector
→ →
ω 4 = ω 3 obtained with the help of the velocity of point C. Velocity of the point E is a
vector perpendicular to DE, in the sense in Figure R 9.7.2 and magnitude
v E = DE ⋅ ω 4 = 2 l ω sin α . Finally, the ICR I 5 of the bar EF is at the intersection of the
→ →
line DE (normal to v E ) with the horizontal of F (normal to v F ). From the isosceles
 ∧   ∧ 
triangle DEF (DE = EF) it is found that mas EDF  = mas DFE  = β and from the right
   

triangle DFI 5 that I 5 F = 2 l sin β and I 5 E = l . The vector v E determines the sense of

the angular velocity ω 5(clockwise) and its magnitude:
v
ω 5 = E = 2 ω sin α .
I 5E
Velocity of the point F (as the velocity of any point of the body 7) is a vector
perpendicular to I 5 F (/ / yy ', oriented towards B) and has the magnitude v F = I 5 F ⋅ ω 5 =
4l ω sin α sin β .

163
Theoretical mechanics – theory and applications

R 9.8) Let us consider the mechanism shown in Figure R 9.8.1 consisting of crank OA and
the disc of center A and radius r. The crank OA rotates about the point O with the angular
velocity ω 0 and angular acceleration ε 0 and the disc rolls without slipping on a cylindrical
surface. Determine the velocities of points A, B and C and acceleration of point A.
Known: OA = 55 cm, r = 20 cm, = 2 rad / s, ε 0 = 5 rad / s.

0
30
r
ω0
O
300 A
ε0
B

Figure R 9.8.1
Solution: The disc has a plane-parallel motion; the ICR is in I 2 (see Figure R
9.8.2). It is easily obtained that:

⊥ OA ⊥ (I BC )
→  → →  2

v A sense given by ω 0 ; ω 2 clockwise sense ;


v = OA ⋅ ω = 110 cm / s  v
 A 0
ω 2 = A = 5,5 rad / s
 I 2A
⊥ I 2 B
→  →
v B sense given by ω 2 ;
v = I B ⋅ ω = 2 r ω sin 150 = 56,94 cm / s
 B 2 2 2

⊥ I 2 C
→  →
v C sense given by ω 2 .
v = I C ⋅ ω = 2 r ω sin 60 0 = 190,52 cm / s
 C 2 2 2

C 
 aA
 a τA
ω2  
300 VC 1
ε0
i2 A
O O
300 ⋅
 A a ν
 ω0 A
 B
VB VA
2

Figure R 9.8.2 Figure R 9.8.3

164
Theoretical mechanics – theory and applications

Acceleration of point A is the acceleration of a point under circular motion (see


Figure R 9.8.3):
→ → →
a A = a τA + a ν
A
⊥ OA // OA
→  → →

a τA sense given by ε 0 ; ν
a A A → O ;
 τ a ν = OA ⋅ ω 2 = 220 cm / s 2
a = OA ⋅ ε 0 = 275 cm / s 2
 A
 A 0

a A= (a τA )2 + (a νA )2 = 352,17 cm / s 2 .
R 9.9) A body rotates about an axis passing through the origin of a Cartesian coordinate
system Oxyz. The velocity of the point M 1 (1,0,1) of the body is equal to v 1= 4 m / s and
the angle α 1 between this velocity and the axis Ox is of 45 0 . Knowing that the angle α 2
between the velocity of the point M 2 (3,4,0) and the axis Ox is given by the
equation cos α 2 = −0,8 , find:
→ →
a) instantaneous angular velocity ω and velocity v 2 of the point M 2 ;
b) equation of the instantaneous axis of rotation.

Solution: a) The velocity of a point M(x, y, z) of a rigid solid with a fixed point is obtained
using the vector relationship:

→ → →
i j k
→ → → → →
v M = ω × OM = ω x ( )
ω y ω z = ω y ⋅ z − ω z ⋅ y i + (ω z ⋅ x − ω x ⋅ z ) j +
x y z

(
+ ω x⋅ y − ω y⋅ x k ) →
(1)
→ →
So that the projections of the velocities v 1 and v 2 are:
v 1 x = ω y ⋅ z 1 − ω z ⋅ y 1= ω y (2a)
v 1 y = ω z ⋅ x 1− ω x ⋅ z 1= ω z − ω x (2b)
v 1 z = ω x ⋅ y 1 − ω y ⋅ x 1 = −ω y (2c)
v 2 x = ω y ⋅ z 2 − ω z ⋅ y 2 = −4ω z (3a)
v 2 y = ω z ⋅ x 2 − ω x ⋅ z 2 = 3ω z (3b)
v 2 z = ω x ⋅ y 2 − ω y ⋅ x 2 = 4ω x − 3ω y (3c)
Denoting by α 1, β 1, γ 1 , and α 2, β 2, γ 2 , respectively the angles made by the vectors
→ →
v 1 and v 2 , respectively with the axes of the system Oxyz, we can write:

165
Theoretical mechanics – theory and applications

v 1x = v 1 cos α 1= 2 2 (4a)
v 1 y = v 1 cos β 1= 3 cos β 1 (4b)
v 1z = v 1 cos γ 1= 3 cos γ 1 (4c)
cos α 1+ cos β 1+ cos γ 1= 1
2 2 2
(4d)
and
v 2 x = v 2 cos α 2 = −0,8 v 2 (5a)
v 2 y = v 2 cos β 2 (5b)
v 2 z = v 2 cos γ 2 (5c)
cos 2 α 2 + cos β 2 + cos 2 γ 2 = 1
2
(5d)

From the relations (2a) and (4a) it follows that:


ω y= 2 2 (rad / s) (6)
and from the relation (2c) it is obtained that:
v 1z = −2 2 (m / s) (7)
Multiplying the relation (4d) with v 12 and using (4a, b, c) we find that:
v 1 y = v 12 − v 1`2x − v 12z = 16 − 8 − 8 = 0 (m / s) (8)
so that from the relations (2b) and (8) it results that:
ω z= ω x (9)
From the relations (3a) and (5a) it is found that:
ω z = 0,2 v 2 (10)
and from (3b, c), (5a), (6), (9) and(10) it results that:
v 2 x = −0,8 v 2 , v 2 y = 0,6 v 2 , v 2 z = 0,8 v 2 − 6 2 (11)

Substituting the relations (11) in


v 22 = v 22 x + v 22 y + v 22 z (12)
we obtain a second degree equation in the unknown solution v 2 that has the double
solution:
15 2
v 2= (m / s) (13)
2
→ →
So that the vectors v 2 and ω are given by:
→ → 9 2 → → 3 2→ → 3 2 →
v 2 = −6 2 i + j , ω= i + 2 2 j+ k (14)
2 2 2

b) The equation of the instantaneous axis of rotation is of the form:


x y z
= = (15)
ωx ωy ωz

166
Theoretical mechanics – theory and applications


Substituting the values of the projections of the vector ω , the instantaneous axis of
rotation is obtained as an intersection of two planes:
3
x= z= y (16)
4

R 9.10) Let us consider a moving cube of side a, for which the velocity v A of the point A,
directed by the diagonal AH, is known at a given time (Figure R 9.10.1). Determine the
velocity of point D knowing that it is directed by the edge AD as well as the velocity of
point B knowing that it is located in the plane ABD.
H G 
VD

E F
 D
 prBD VD 
 VD VB
VA
D C

VB
A 
A B prBD V B
B

Figure R 9.10.1 Figure R 9.10.2

Solution: We know that in the general movement of the solid rigid the projections
of the velocities of two points belonging to the rigid on the line connecting the two points
are equal.
Using this property for points A and D it is found that:
→ →
pr AD v D = pr AD v A (1)
Velocity of the point D, directed as assumed, by the edge AD, shall have,

therefore, the sense of the projection of v A on AD (namely from A to D) and the
magnitude:
→ 2
v D = pr AD v A = v A cos 45 0 = v (2)
2 A
Applying the same property for the points A and B we find that:
→ →
pr AB v A = pr AB v B= 0 (3)
→ →
since v A ⊥AB , which means also that v B ⊥AB .
Using also the assumption that the velocity of the point B is located in the plane

ABD it follows that the vector v B is directed by the edge BC. In order to obtain the scalar
of this velocity, notice that:
→ →
pr BD v D = pr BD v B (4)

167
Theoretical mechanics – theory and applications

Since they are parallel vectors and have the same projection on the line BD (Figure
R 9.10.2) it results that:
→ →
v B= v D (5)
9.5. Problems proposed

9.5.1. Conventional tests

TC 9.1) The wheel of a swing is rotating at a constant speed n = 60 rot / min. Knowing
that the radius of the wheel is R = 10 m, determine the velocity and acceleration of point B
of a seat if the seat axis remains parallel to itself throughout the motion (Figure TC 9.1.1).

TC 9.2) In the rotation of a grinder disc it is known that at an instant the size of the angle
at the center described from the beginning of the motion is proportional to the square of
time. Determine the velocity and acceleration of any point on the periphery of the disk of
radius R = 30 cm at time t 2 = = 4 s if after t 1 = 2 s from the beginning of the motion the
disc has the speed n 1 = 50 rot / min.
n 
V1
A A
B
O
R

O


B V2

Figure TC 9.1.1 Figure TC 9.3.1


TC 9.3) Two stiff-united wheels of radius R and r, respectively, are driven simultaneously
→ →
with two racks so that the points A and B have known velocities v 1 and v 2 , respectively.
If v 1 > v 2 , determine the position of the instantaneous center of rotation, instantaneous
angular velocity and velocity of the common center O of the two wheels (Figure R 9.3.1).


ω1
M
• 2α

M0 2β

Figure TC 9.4.1

168
Theoretical mechanics – theory and applications

TC 9.4) Let us consider two right circular cones of equal length slants l = 30 cm, angles at
vertex 2α = 90 0 and 2 β = 60 0 , respectively (Figure TC 9.4.1). The first cone is fixed and
the second one rolls over the fixed one at the angular velocity ω 1= 1,2 rad / s directed
along the symmetry axis of the fixed cone. The two cones have the vertex at the same point
O. Determine the angular velocity and acceleration of the mobile cone and the velocity and
acceleration of point M which finds on the line of intersection between the mobile cone
base plane and the plane formed by the two axes of symmetry of the cones, at the distance
MM 0 = 10 cm from the circumference of the base circle.
TC 9.5) A rigid solid (C) is considered in the general motion studied with the help of the
fixed reference system O 1 x 1 y 1 z 1 and movable reference system Oxyz. We know at an
instant the velocities:
→ → → → → → → → → → →
v O = 3 i − 2 j , v A= 2 i + 4 j − 4 k , v B = 5 i + 6 j − 2 k
of three points O (0, 0, 0), A (2, 1, 1) and B (2, 0, 2) belonging to the rigid solid (Figure
TC 9.5).
→ →
a) determine the elements ω and v min of the kinematic torsor of reduction relative to one
point of the instantaneous axis of the helical motion;
b) equations of the instantaneous axis of helical motion.
AIME
z
z

ω  y
B(2,0,2)  Vmin
VB

rB
A(2,4,4)

V0
O′ 
VA
 ( C)
ω
x u= y
O ω

Figure TC 9.5

9.6. Indications and answers


→ →
TC 9.1) See Figure TC 9.1.2. The chair has a translation motion, hence v B = v A
→ →
and a B = a A.

⊥ OA  // OA
 → → 

 →
ν
v A motion sense (sense given by ω ) , a A= a A A → O
v = OA ⋅ ω = 20 π (m / s )  ν
 a A = OA ⋅ ω = 40 π
2 2

A (m / s 2 )

169
Theoretical mechanics – theory and applications

The relation between wheel speed (expressed in rot/min) and its angular velocity
πn
(expressed in rad/s): ω = = 2 π has been used).
30

• •• πn
TC 9.2) θ = λ t 2 ( λ > 0 constant), ω = θ = 2 λ t , ε = θ = 2 λ . Since ω = , it follows
30
π n1 π n2 t2
that = 2 λt 1, = 2 λ t 2 and hence, n 2 = n = 100 rot / min .
30 30 t1 1
π n2 v 22 250 π 2
v 2= R ω 2= R ⋅ = 100 π cm / s , a ν2 = = cm / s 2 ,
30 R 3
π n1 π
a τ2 = R ε 2 = R ⋅ 2 λ = R ⋅
30 t 1
=
4
cm / s 2 , a 2 = (a ν2 )2 + (a τ2 )2 .

A V1 A′
A
  
aA V0
VA O
V2
O B B B′

aB 
VB

ω
I

Figure TC 9.1.2 Figure TC 9.3.2

TC 9.3) See Figure TC 9.3.2. Let I be the instantaneous center of rotation.


→ →
Since v 1 ⊥ IA , v 2 ⊥ IB , it results that I ∈ AB . Velocity distribution on IA is linear,
hence I ∈ A ' B ' . It results that I finds at the intersection of the lines AB and A ‘B ‘.
IB v2 v2
∆ IBB ' ≈ ∆ IAA ' ⇒ = ⇒ IB = ( R + r ) .
IB + R + r v 1 v 1− v 2

Angular velocity ω is perpendicular to the figure plane, has a clockwise sense and
v 2 v 1− v 2
the magnitude ω = = . Velocity of the center O is perpendicular to IO, is
IB R+r
v1 r + v 2R
oriented to the right and has the magnitude v O = IO ⋅ ω = ( IB + r ) ⋅ ω = .
r+R
TC 9.4) The mobile cone motion is the motion of a rigid with fixed point (point O). In this

motion the angular velocity vector ω is directed along the slant OO '(axis of zero velocity

points at an instant) and it is a resultant of the composition of two vectors: ω 1 -

characterizes the rolling motion of the movable cone over the fixed one and ω 2 -

170
Theoretical mechanics – theory and applications

characterizes motion of the movable cone around its own axis. Applying the sines theorem
in the triangle formed by the angular velocity vectors ( ∆OAB ) we find that:
ω1 ω2 ω
= = ,
sin 30 0 sin 45 0 sin 105 0
sin 105 0
hence ω = ω 1 = 2,32 rad / s . Velocity of the point M is given by:
sin 30 0
 ∧ 
→ → → → → → → 
v M = ω × OM and v M = ω × OM = ω ⋅ OM ⋅ sin ω , OM  .
 
 
OM O'M
In ∆OO ' M we apply the sines theorem and obtain that = ,
sin 60 0 sin 105 0
hence v M = ω ⋅ O ' M ⋅ sin 60 0 = 40,14 m / s (since the radius of the cone of angle at vertex
2 β is R 2 = l sin β = 15 cm and therefore O’M = 20 cm).
Considering a fixed Cartesian reference system O x 1 y 1 z 1 with the axis O z 1

along the vector ω 1 , we have that:
→ → → →
ω = ω sin 45 0 cos ω 1 t i 1 + ω sin 45 0 sin ω 1t j 1+ ω cos 45 0 k 1 ,
hence:

→ → → →
ε = ω = −ω ω 1sin 45 0 sin ω 1t i 1 + ω ω 1 sin 45 0 cos ω 1t j 1
→ →
and, for t = 0, ε = ω ω 1sin 45 0 j 1 and ε = ω ω 1sin 45 0 = 1,96 rad / s 2 . Eventually,
→ → → → → → 
a M = ε × OM + ω ×  ω × OM  .
 
→ → →
Remarking that at t = 0 OM = OM sin( 45 + γ ) i 1 + OM cos(45 + γ ) j 1 and using
→ →
the expressions of ω and ε at t = 0 we obtain the acceleration a M = 64,2 cm / s 2 . γ
→ →
denotes the measure of the angle between the vectors ω and OM .

TC 9.5) a) Components of the velocity of an arbitrary point M(x, y, z) of a rigid in general


motion are:
v x = v O x + ω y⋅ z − ω z⋅ y , v y = v O y + ω z⋅ x − ω x⋅ z , v z = v O z + ω x⋅ y − ω y⋅ x .

Apply this result for the points A and O, and B and O, respectively, and find
that ω x = −2 , ω y = 1 , ω z = 2 , hence:
→ → → → →
ω = −2 i + j + 2 k , ω = ( − 2) 2 + 1 2 + 2 2 = 3

171
Theoretical mechanics – theory and applications

→ →
→ ω⋅v → 16 → 8 → 16 →
⋅ω =
A
v min = i − j− k.
ω 2 9 9 9

b) The instantaneous axis of the helical motion is given by the equations:


v Ox− y ω z+ z ω y v O y− z ω x+ x ω z v Oz− x ω y+ y ω x
= =
ωx ωy ωz
In the particular case of our problem it is obtained that:
8 11
x+2 y= , 2y − z = .
9 9

172
Theoretical mechanics – theory and applications

10. Relative motion of the material point


In Chapters 8 and 9 we studied the motion of a material point or a rigid with
respect to a fixed reference (i.e. absolute motion).
We shall further consider the motion of a material point (Chapter 10) in relation to
a movable reference, found in its turn, in motion with respect to a fixed reference. The
question in this case, is the problem of determining the kinematic parameters (trajectory,
velocity, acceleration) characterizing the motion of the point or rigid relative to fixed
reference, if the kinematic parameters characterizing the motion of the point or rigid
relative to the movable reference and kinematic parameters characterizing the motion of
the movable reference relative to the fixed one are known.

10.1. Definitions and examples

The material point motion in relation to the fixed system is called absolute motion.
Velocity (or acceleration) of the point in this motion is called absolute velocity (absolute
→ →
acceleration, respectively) and is denoted by v a ( a a , respectively).
The material point motion in relation to the movable system is called relative
motion. Velocity (or acceleration) of the point in this motion is called relative velocity
→ →
(relative acceleration, respectively) and is denoted by v r ( a r , respectively).
It is called transport motion, the motion of the point stiff-connected to the movable
reference in relation with the fixed system and which at the considered instant coincides
with the point whose motion is studied. Velocity (acceleration, respectively) of the point in
this motion is called transport velocity (transport acceleration, respectively) and is
→ →
denoted by v t ( a t , respectively).

Example: Let a bar OA be in rotation motion about the joined end at O in the fixed
plane O 1 x 1 y 1 (Figure T 10.1). While the bar rotates, a material point M (a slider) slides
along the bar from O to A. The absolute motion is the motion of the point relative to the
fixed reference O 1 x 1 y 1 . With the point moving along OA (considered as axis Ox of the
movable reference Oxy) and the bar rotating about OA the trajectory of the point shall be a
spiral type curve. The relative motion is the motion of the point with respect to the
movable reference Oxy, hence a rectilinear motion on OA. The transport motion is the
motion of the point M considered immovable with respect to the movable reference and in
motion with respect to the fixed one. It is a circular motion on the circle with center at O
and of radius OM.

173
Theoretical mechanics – theory and applications

y1 x z1 z
y y
A M r
M
 •
r1
 O y1
rO
O1 x
x1
O1 = O
x1

Figure T 10.1 Figure T 10.2

10.2. Absolute derivative and relative derivative of a vector



Let the vector v (t ) be defined by its projections on the axes of the movable
reference Oxyz:
→ → → →
v (t ) = v x i + v y j + v z k (10.1)

Differentiating the relation (10.1) with respect to time t we obtain:



d v →  • → • → • →   → →
• • • •

= v =  v x i + v y j + v z k  + vx i + v y j + v z k  (10.2)
dt    
 
The left member of the relation (10.2) represents the total or absolute derivative of
→ →
the vector v . The first bracket in the right member represents the vector v derivative
→ → →
with respect to the movable system as if it would be fixed (that is versors i , j and k do
not change their direction). This derivative is called the local or relative derivative of the

→ ∂v
vector v and is denoted by , with the mention that this notation does not represent a
∂t
partial derivative. Taking into account the Poisson equations:
• • •
→ → → → → → → → →
i = ω× i , j = ω× j , k = ω×k
The second bracket takes the form:
• • •
→ → →  →
→ → → → →
v x i + v y j + v z k = ω×  v x i + v y j + v z k  = ω× v (10.3)
 
 
Relation (10.2) becomes:
→ →
dv ∂v → →
= + ω× v (10.4)
dt ∂t

This relation represents the formula for obtaining the absolute derivative (with

respect to the fixed reference) of a vector v given by its projections on the movable
reference axis.

174
Theoretical mechanics – theory and applications

10.3. Composition of velocities

Let a material point M be in absolute motion relative to the fixed reference


O 1 x 1 y 1 z 1 and the movable reference Oxyz (Figure T 10.2). The position of the point

relative to the fixed reference is given by the position vector r 1 and relative to the movable

reference by the vector r . The relation between the two vectors is:
→ → →
r1 = r O + r (10.5)

where r O is the position vector of the origin of the movable trihedral relative to the fixed
one.
Differentiating the relation (10.5) with respect to time we obtain:
• • •
→ → →
r 1 = r O+ r (10.6)
But:

→ →
r 1 = v a - is the absolute velocity (with respect to the fixed reference);

→ →
r O = v O - is the absolute velocity of the origin O of the movable reference;
• →
→ ∂ r → →
r = + ω× r
∂t

∂ r →
= v r - is the relative velocity (with respect to the movable reference, as this one were
∂t
fixed).
→ → →
Noting that v O + ω× r represents the velocity of a point stiff-connected to the
movable trihedral coinciding at the considered instant with the point M, that is the

transport velocity v t , we obtain the following formula for composition of velocities in the
relative motion of the material point:
→ → →
va = vr+ v t (10.7)

10.4. Composition of accelerations

Let us consider the relation (10.7) written as:



→ ∂ r → → →
va = + v O + ω× r
∂t
which we differentiate with respect to time t:
•  → • • •
→ d ∂ r  → → → → →
va =   + v O + ω× r + ω× r (10.8)
dt  ∂ t 
 

175
Theoretical mechanics – theory and applications

But:

→ →
v a = a a - is the absolute acceleration of the point M;

→ →
v O = a O - is the absolute acceleration of the point O;
• • →
→ → →
∂ r → → → → →
ω = ε ; r = + ω× r = v r + ω× r ;
∂t
 → → → →
d  ∂ r  ∂2 r → ∂ r ∂2 r → →
 = + ω× = + ω× v r ;
dt  ∂ t  ∂ t 2 ∂t ∂ t2
 

∂2 r →
= a r - is the relative acceleration of the point M.
∂ t2

→ → → → → →
Sum a O + ε × r + ω ×  ω × r  represents acceleration of a point stiff-connected to
 
 
the movable trihedral coinciding at the considered instant with the point M, that is the
→ → → not →
transport acceleration a t of the point M. The expression 2 ω × v r = a C is called Coriolis
acceleration and expresses the simultaneous influence of the rotation motion of the
movable trihedral relative to the fixed one and of the relative motion of the point on the
absolute acceleration of the material point. Taking into account these observations we
obtain the following formula for the composition of the accelerations in the relative motion
of the material point:
→ → → →
aa = ar+ a t+ aC (10.9)

10.5. Solved problems

 ∧

R 10.1) Slider M moves along the bent bar OAB ( m OAB  = 90 , OA = 12 cm)
 
following the law AM = x (t) = 4 t 2 + t (Figure R 10.1.1). The bar rotates in its plane
about an axis passing through O, at the angular velocity ω = 3 t . Determine the absolute
velocity and the absolute acceleration of the slider M depending on time. Customize the
results found for t 1 = 1 s.
Solution: Let us consider the fixed reference system O1 x1 y1 z1 and the movable
reference system Axyz, with respect to which we shall study the motion of the slider M.
The relative motion of the point M is a rectilinear motion on AB, in compliance with
the law x = 4 t 2 + t . The transport motion is obtained by joining the point to the bar (that is
making the relative motion to cease).

176
Theoretical mechanics – theory and applications

A x

M B

ω
O

Figure R 10.1.1

Point M shall make in this case a circular motion on the circle with center at O and of
radius AM = l 2 + x 2 , at the angular velocity ω = 3 t .

Study of velocities (Figure R 10.1.2)


→ → →
va = vr + vt
// AB ⊥ OM
 → 

 →
v r A → B ; v t sense given by ω
 
v t = OM ⋅ ω = 3 t l + (4 t + t )
. 2 2 2
v r = x = 8 t + 1

A x A x
y1
⋅ 
y1
⋅ 
• M
Vr ⋅ •M ar
• B B

x a νt 
a τt x

Vt 
Va 
 at
θ θ ac
α α

 
O1 ω x1 O1 ε x1

ω

y y

Figure R 10.1.2 Figure R 10.1.3

v a = v 2r + v t2 + 2 v r v t cos α = (8 t + 1) 2 + 9 t 2[144 + (4 t 2 + t )2 ] + 72 t (8 t + 1)

177
Theoretical mechanics – theory and applications

Study of accelerations (Figure R 10.1.3)


→ → → → →  // AB → → →

aa = a r+ a t+ aC ; a r A → B ; a t = a τt + a νt
 ..
a r = x = 8

⊥ OM // OM
→  
 → →
a τt sense given by ε ; a νt M → O
 τ  ν
a t = OM ⋅ ε = OM ⋅ ω = 3 l 2 + (4t 2 + t ) 2 a t = OM ⋅ ω 2 = 9t 2 l 2 + (4t 2 + t ) 2
.

 → → 
⊥  ω , v r 
→ → → 
a C = 2ω × v r sense given by drill rule

aC = 2 ω v r sin  ω , v r  = 6 t (8 t + 1)
→ →

  

a a x = a r + a τt cos α − a νt sin α = −36 t 4 − 9 t 3+ 44

a a y = a C + a τt sin α + a νt cos α =`168 t 2 + 9 t

a a = a 2a x + a 2a y

Particular case: t = 1 s
12 5
x = 5 cm , OM = 13 cm , cos α = , sin α = , ω = 3 rad / s , ε = 3 rad / s 2
13 13
v r = 9 cm / s , v t = 39 cm / s , v a = 43,4 cm / s
a r = 8 cm / s 2 , a τt = 39 cm / s 2 , a νt = 117 cm / s 2 , a C = 56 cm / s 2 , a a = 177,1 cm / s 2

ω 02
R 10.2) A slider M moves on the semicircle of radius R as per the law φ 1= t 2 , where
4
ω 0 is a positive constant. At the same time, the semicircle rotates about its diameter AB
as per the law AB φ 2 = ω 0 t (Figure R 10.2.1). Determine:
π2
a) Time t 1 after which the slider attains the position φ 1= ;
4
b) For the position corresponding to the instant t 1 , determine the absolute velocity
and acceleration of the slider M.
178
Theoretical mechanics – theory and applications

B
ϕ2

O•

ϕ1

M
A

Figure R 10.2.1

π 2
ω 2
π π
Solution: a) ϕ 1(t 1 ) =
0
⇒ t 12 = ⇒ t 1=
4 4 4 4

b) The relative motion is a circular motion on the circle with center at O and of radius
ω2
OM=R, as per the law ϕ 1(t ) = 0 t 2 .
4
The transport motion is equally a circular motion, on the circle of center N and
ω 2 
radius MN = OM sin ϕ 1= R sin 0 t 2  , as per law ϕ 2 (t ) = ω 0 t (Figure R 10.2.2).
 4 
 

Study of velocities (Figure R 10.2.2)

→ → → y
  v a= v r+ v t 
B Vr Vt B a τr
⊗ x •
z
M • a
 
N M N at  c
a νr
O
O 
 ε1
ω1 ϕ1
 ϕ1
ω2
A
A

Figure R 10.2.2 Figure R 10.2.3

179
Theoretical mechanics – theory and applications


⊥ (OAM ) // AB
→ 
 → 
ω 1  trigonometric sense ; ω 2  A → B
 
ω 1= ϕ 1= ω 0 t
. 2 .
ω 2 = ϕ 2 = ω 0
 2
 
⊥ OM ⊥ (OAM )
→  → 
 →
 →
v r sense given by ω 1 ; v t sense given by ω 2
 
v r = OM ⋅ ω 1= π Rω 0 v t = MN ⋅ ω 2 = Rω 0 sin π
2 2

 4  4

→ → π 2ω 2
π 2
v r ⊥ v t ⇒ v a = v 2r + v t2 = R ω 0
0
+ sin 2
64 4

Study of accelerations (Figure R 10.2.3)

→ → → →
a a= a r+ a t+ a C

⊥ (OAM )
→  → → → → → → →
 → → →
ε 1  trigonometric sense ; ε 2 = ω 2 = 0 ; a r = a τr + a νr ; a t = a νt  a τt = 0 
  
ε 1= ω 1= ω 0
. 2

 2

  
⊥ OM // OM // MN
→   → 
 → →
a τr sense given by ε 1 ; a νr M → O
ν
; a t M → N
  
a τr = OM ⋅ ε 1= Rω 0 a νr = R ω 12 = π Rω 0 a νt = MN ⋅ ω 22 = R ω 02 sin π
2 2 4 2

 2  64  4

 → → 
⊥  ω 2 , v r 
  
→ → →

a C = 2ω 2 × v r  trigonomet ric sense

a C = 2ω 2 v r sin  ϕ 1− π  = π R ω 0 sin  π − π 
3 2

  4 2 
 2 4 

a a x = a tν + a τr cos(π − ϕ 1 ) + a νr sin(π − ϕ 1 ) ; a a y = a τr sin(π − ϕ 1 ) − a νr cos(π − ϕ 1 )

180
Theoretical mechanics – theory and applications

a a z = a C ; a a = a 2a x + a 2a y + a 2a z

R 10.3) A circular disc of radius R = 2r, with the center at O, rolls without slipping on the
line (D). In a circular groove of radius r, centered at O, a movable item M is moving at the

constant velocity of magnitude v 1 = u = constant (Figure R 10.3.1). Determine the

absolute velocity and acceleration of the movable item when it finds at the position M 0
(on the diameter parallel to the line (D)) knowing that the center O of the disc is moving at

the constant velocity v O = 2u (oriented to the right).

M



V1

O V0
r M0

( D)

Figure R 10.3.1

Solution: The relative motion is a circular motion on the circle centered at O and of

radius r, at the velocity v 1 . The transport motion is the motion of a point (M 0 ) of a body
in plane-parallel motion.

Study of velocities (Figure R 10.3.2)

The ICR of the disc is at the point I of contact with the line (D)


⊥ (IOM 0 ) ⊥ OM 0
  ⊥ OM 0

  →
 → → →

ω clockwise sense  given by v 0  ; v t sense given by ω ; v r downwords
    v = v = u
 v0 u v t = IM 0 ⋅ ω = u 5

 r 1
ω = =
 r r

→ → → → → 
v a= v r+ v t ; v a = v 2r + v t2 + 2 v r v t cos v r , v t  = 2 2 u
 

181
Theoretical mechanics – theory and applications

 
ω


ac

V0
O M0 O J = 0 a t M0 ⋅
•     •
Vt aa ac ar

Vr  
Va Vr

 
I ω I ω

Figure R 10.3.2 Figure R 10.3.3 Figure R 10.3.4

Study of accelerations (Figure R 10.3.3)


→ → →
v O = cons tan t ⇒ a O = 0 ⇒ the acceleration pole for the disc is at the point O ( J ≡ O ) .
→ → →
ω = cons tan t ⇒ ε = 0 ;
 
 // OM  // OM 0
→ →  0 → → 
τ  ν 
a r = a r M 0 → O ; a t = a t M 0 → O
 2 2  2
a = v 1 = u a = OM ⋅ ω 2 = u
  t
r 0
r r r
  → →

⊥  ω , v r 
→ 
 
→ →

a C = 2 ω × v r sense given by drill rule (see figure R 8.3 c )

a C = 2 ω v r sin  ω , v r  = 2 u
→ → 2

   r
→ → → →
a a= a r+ a t+ a C
2
4u
a a= a r+ a t + a C = .
r

R 10.4) The ladder AB, of length l, moves so that the end A moves at the velocity

v A = u on a horizontal wall and the end B on a vertical wall (Figure P 10.4.1). On the

bar, a material point M is moving from B to A as per the law s (t ) = v ⋅ t (u, v are
constantly positive). Requested:
→ → → →
a) v B , a B , ω and ε , respectively, in the plane-parallel motion of the bar;
b) Absolute velocity and absolute acceleration of the point M at the
π l
position θ = , s = .
4 2

182
Theoretical mechanics – theory and applications

B
S

θ
M


VA
A

Figure R 10.4.1

Solution: a) The position of the bar at an instant is given by the angle θ ( t ) (see
Figure R 10.4.2). The ICR of the bar AB finds at the intersection of the perpendicular at A
→ →
on Ox ( ⊥ v A ) with the perpendicular at B on Oy ( ⊥ v B ), that is it is the fourth corner of
the rectangle BOAI .


⊥ (OAB )
 ⊥ IB
→  → 

ω  trigonometric sense  given by v A  ; v B sense given by ω (downwards )
→ →

   v = IB ⋅ ω = u tg θ
 vA u  B
ω = =
 IA l cosθ

⊥ (OAB )


 .
u sin θ . u 2 sin θ .
ε ε = ω = θ= 2 , as ω = θ .
 l cos θ
2
l cos θ 3

 trigonometric sense (ε , ω > 0 )

→ → →
v A = cons tan t ⇒ a A = 0 ⇒ A ≡ J (acceleration pole)
 
⊥ JB // JB
→ → → →  → → 
a B = a τB + a νB ; a τB sense given by ε ; a νB B → J
 
a τB = JB ⋅ ε = u sin θ
2 2
a νB = JB ⋅ ω 2 = u 1
 l cos 2θ  l cos 2θ

a B= (a τ ) + (a ν )
B
2
B
2
=
u2 1
l cos 3θ
(// BO )

183
Theoretical mechanics – theory and applications

→ .
Or else: Point B has a rectilinear motion, hence a B // BO , B → O , a B = v B .

b) The relative motion is a rectilinear motion on BA as per the law s (t) = v t. The transport
motion is the motion of a point M of a body (bar AB) in plane-parallel motion.

y y

B i
 B
a τB  I 
 a νB ω π/4
aB π/4
 θ
VB l M ⋅ 
Vt

Vr
 
VA
Va
O O
A≡J x A x

ε

Figure R 10.4.2 Figure R 10.4.3

Study of velocities (Figure R 10.4.3)



// BA ⊥ IM (// BA)
→ 

 →
 →
v r B → A ; v t sense given by ω (towards A)
 . 
v r = s = v = cons tan t v = IM ⋅ ω = l u
=
u 2
 t 2 l cos π 2
 4

y ac
B I
  

aB π / 4
ac ω ω • ⋅

M aa 
Vr

at
O
A≡J x

Figure R 10.4.4 Figure R 10.4.5

184
Theoretical mechanics – theory and applications

→ → → u 2
v a= v r+ v t ; va = v r + v t = v +
2

Study of accelerations (Figure R 10.4.4)


 // OB
→ → → 1→ 
ar = 0 ; a t= aB B → O
2  2
a t = u 2
 l
  → →
⊥  ω , v 
→ → →   
a C = 2ω × v r sense given by drill rule (see Figure R 10.4.5)

a = 2 ω v sin  ω , v  = 2 u v 2
→ →

 C r

r
 l

→ → → → → →  u
a a= a r+ a t+ a C ; a a = a t2 + a C2 + 2 a t a C cos a t , a C  = 2u2 + 8v 2− 4 2 u v
  l

10.6. Problems proposed

10.6.1. Conventional tests

TC 10.1) Be it the mechanism in Figure TC 10.1.1, formed by the bars O 1 A and O 2 B of


length R and the semi disc centered at C and of radius AC = CB = R. Bar O 1 A rotates
about the point O 1 at the constant angular velocity ω . On the semi disc a material point
 ∧  ε t2
M moves so that m BCM  = θ ( t ) = , where ω and ε are constant.
  2
y
A
O1 O2
1  
ω y = x2 V2 V1
6

C B O
A x

Figure TC 10.1.1 Figure TC 10.2.1

185
Theoretical mechanics – theory and applications

Determine the absolute velocity and acceleration of the point M at an arbitrary instant t.


TC 10.2) A trolley is moving on a rectilinear road at the constant velocity v 1 = u . On the

1 2
trolley there is mounted a tube OA with the shape of a parabola of equation y =
x
2
(Figure TC 10.2.1). Inside the tube a material point M moves at the constant

velocity v 2 = 2 u . Determine the absolute velocity and acceleration of the point M at the

instant it passes through the point of abscissa x = 3.

10.6.2. Multiple-choice tests

TG 10.1) A disk of radius R rotates with the constant angular velocity ω around an axis
passing through its center and perpendicular to the disk plane (Figure TG 10.1). On a
diameter of the disk, a point M moves starting from its center as per law s = R sin ω t .
The absolute trajectory of the point M is:
a) Ellipse of centre O and semi axes R and R / 2 directed along the diameter OM and
perpendicular on it; b) A parabola of centre O and diameter OM as axis of symmetry; c)
Circle of center (O, R/2) and radius R / 2; d) Circle of center O and radius R / 2.

M
s •M

 O
ω 
µ ω

Figure TG 10.1 Figure TG 10.2

TG 10.2) A point M starts from the vertex O of a cone with the angle at vertex 2 α and

moves on a slant of the cone at the constant velocity u . In the same time, the cone rotates

about its axis of symmetry at the constant angular velocity ω (Figure TG 10.2). What is
the absolute velocity of the point M after t seconds from the start of the motion?
a) v a = u 1 + ω t sin α ; b) v a = u 1 + ω 2 t 2 sin α ; c) v a = u + ω t sin α ;

d) v a = u 1 − ω 2 t 2 sin α .

186
Theoretical mechanics – theory and applications

10.7. Indications and answers

TC 10.1) The relative motion is a circular motion on the semi disc of diameter AB by
1
law θ ( t ) = ε t 2 . Since O1 ABO2 is a parallelogram it follows that AB remains parallel
2
to itself during motion, thus the semi disc has a translation motion. The transport motion is
the motion of point M considered as a fixed point of the semi disc. The position of the bar
O 1 A is given by the angle ϕ = ω t .

Study of velocities (Figure TC 10.1.2)

⊥ CM ⊥ O1 A
→  → 
 →
 →
v r towards increase of angle θ ; v t = v A sense given by ω
 . v = O A ⋅ ω = R ω
vr = CM ⋅ θ = R ε t 
t 1

→ → →
v a= v r + v t ; v a = v 2r + v t2+ 2v r v t cos(θ + ϕ )

O1 O2
O1 O2
 ϕ
ϕ ω

ω


VA
aA

ν

A ⋅
C
B A
B

Vr
⋅ 
θ C θ
ac
 
Vr at

π a νr ϕ
M ϕ−
⋅ 2 M 
π/2+θ  aC
 τ a τr
Vr

Figure TC 10.1.2 Figure TC 10.1.3 Figure TC 10.1.4

Study of accelerations (Figure TC 10.1.3)

⊥ CM 
 // CM
→ → → → →

a r = a τr + a νr ; a τr towards increase of angle θ ; a νr M → C
 τ ..  . 2
a r = CM ⋅ θ = R ε a νr = CM ⋅ θ = Rε 2 t 2

187
Theoretical mechanics – theory and applications


// O A
→ → → 1
ν 
a t = a A = a A A → O1 ;
 ν
 a A = O1 A ⋅ ω = R ω
2 2

 → → 
⊥  ω , v r 
→ → →

a C = 2 ω × v r sense given by drill rule (Figure TC 10.1.4);

a C = 2 ω v r sin  ω , v r  = 2 R ω ε t
→ →

  
→ → → →
a a= a r+ a t+ a C

a τa = a τr + a t sin( ϕ + θ − π ) ; a νa = a C − a νr − a t cos( ϕ + θ − π ) ; a a = (a τa )2 + (a aν )2
TC 10.2) The relative motion is the motion of the point M related to the tube at the
constant velocity of magnitude v 2 = 2u . The transport motion is the motion of the point M
considered as a point of the trolley in translation motion at the constant speed v 1 = u .

Study of velocities (Figure TC 10.2.2)

The angle made by the tangent at M to the parabola with the axis Ox is given by the
π 1
relation α = arctg f ' ( 3) = , where f ( x) = y = x 2 .
4 6
on tan gent in M to the parabola // Ox

 → →

v r O → A ; v t = v 1 to right
v = v = 2 u v = v = u
 r 2  t 1

→ → →
v a= v r+ v t ; v a = v r2 + v t2 + 2v r v t cos α = u 5 + 2 2

y y
A A
  
Vr   ρ(3) C a r = a 0
Va V1
M
M

Vt

O
O 3 x
3 x

Figure TC 10.2.2 Figure TC 10.2.3

188
Theoretical mechanics – theory and applications

Study of accelerations (Figure TC 10.2.3)

The radius of curvature of the trajectory is given by the relation


 x 2 1,5
1 + 

ρ ( x) =
[ ]
1 + ( f ' ( x) )2 1,5 
=
 9 
.
f ' ' ( x) 1
3


on normal in M to the parabola
→  →
ν 
→ → → →
a r = a r towards centre of curvature ; a t = a 1= 0 (as v 1 = constant )
 2
a ν = v 2 = 2 u 2
 r ρ (3) 3
→ →
a C = 0 (the transport motion is a translation)
→ → → → 2 2
a a= a r + a t + a C ; a a= a r = u .
3

TG 10.1) The absolute coordinates of point M with respect to the Cartesian reference
O x 1 y 1 with axes O x 1 and O y 1 horizontal and vertical, respectively, are:
R
x 1= s cos ω t = R sin ω t cos ω t = sin 2 ω t
2
y 1= s sin ω t = R sin 2 ω t = (1 − cos 2 ω t ) .
R
2
By eliminating the time variable ( sin 2 ω t + cos 2 2 ω t = 1 ) the relation
2

 R 2
x 12 +  y 1− = 1 is found representing the equation of the circle with center at (O, R/2)
 2
and radius R / 2. Correct answer: c).

TG 10.2) The relative motion is the motion of the point M on the slant at the velocity

u and the transport motion is the circular motion on the circle of center M ‘ and radius
MM ‘ (where M‘ is the projection of the point M on the axis of symmetry of the cone) at

the angular velocity ω . Hence, v r = u and v t = MM '⋅ω = OM sin α ⋅ ω = u t sin α ⋅ ω .
→ → → → →
Since v r ⊥ v t and v a = v r + v t , we find that v a = v r2 + v t2 = u 1 + ω 2 t 2 sin 2 α .
Correct answer: b).

189
Theoretical mechanics – theory and applications

11. Relative motion of the rigid solid


11.1. Generals
Let a rigid (C) be in motion relative to the fixed reference O x 0 y 0 z 0 and the
movable reference O 1 x 1 y 1 z 1 . We denote by O 2 x 2 y 2 z 2 a movable trihedral stiff-
connected to the rigid (Figure T11.1). We intend to determine the kinematic parameters
that characterize the rigid motion relative to the fixed reference when we know the rigid
motion with respect to movable trihedral O 1 x 1 y 1 z 1 and the motion of this movable
trihedral with respect to the fixed one.
z0 z2


( C)
M
z1 O2 y2

x2 y1
O1
y0
O
x1

x0

Figure T 11.1
Motion of the trihedral O 2 x 2 y 2 z 2 related to O 1 x 1 y 1 z 1 is defined by the
vectors:

v 21 - velocity of origin O 2 related to O 1 x 1 y 1 z 1 ;

ω 21 - angular velocity in the relative motion of the rigid related to O 1 x 1 y 1 z 1 ;

a 21 - acceleration of the origin O 2 related to O 1 x 1 y 1 z 1 ;

ε 21 - angular acceleration in the relative motion of the rigid related to O 1 x 1 y 1 z 1 ;

Motion of the trihedral O 1 x 1 y 1 z 1 related to the fixed trihedral O x 0 y 0 z 0 is


defined by the vectors:

v 10 - velocity of origin O 1 related to the fixed trihedral;

ω 10 - angular velocity in the motion of the trihedral O 1 x 1 y 1 z 1 related to the fixed
one;

a 10 - acceleration of origin O 1 related to the fixed trihedral;

190
Theoretical mechanics – theory and applications


ε 10 - angular acceleration in the motion of the movable trihedral O 1 x 1 y 1 z 1 related to
the fixed one.

To obtain the distribution of speeds and accelerations in relative motion of the


rigid it is needed to determine the absolute velocity and absolute acceleration of an
arbitrary point M of rigid as well as the absolute angular velocity and absolute angular
acceleration vectors in the rigid motion with respect to the fixed trihedral.

11.2. Composition of velocities

Let us consider an arbitrary point M of the rigid (Figure T 11.1). Its motion relative
to the trihedral O 1 x 1 y 1 z 1 is a relative motion so that:
→ → → →
v r = v 21 + ω 21 × O 2 M (11.1)
The transport velocity shall be the velocity related to the fixed trihedral of a point
stiff-connected with the trihedral O 1 x 1 y 1 z 1 and that will coincide with the point M at the
considered instant, hence:
→ → → →
v t = v 10 + ω 10 × O 1 M (11.2)
The absolute velocity of the point M will be:
→ → → → → → → → →
v aM = v r + v t = v 10 + v 21 + ω 10 × O 1 M + ω 21× O 2 M (11.3)
Generalizing for more transport motions we obtain:
→ n → n → →
v a =
M
∑ v i,i−1 + ∑ ω i,i−1 × O i M (11.4)
i =1 i =1

In order to determine the absolute angular velocity ω n 0 in the motion of the
trihedral O n x n y n z n , stiff-connected to the rigid with respect to the fixed trihedral we
shall consider another point of the rigid (point N) and rewrite the relation (11.4) for that
point:
→ n → n → →
v a =
N
∑ v i,i−1 + ∑ ω i,i−1 × O i N (11.5)
i =1 i =1
Selecting point M as origin of the trihedral stiff-connected to the rigid, Euler’s
equation for velocities is written:
→ → → →
v a =
N
v a +
M
ω n 0 × MN (11.6)
→ → →
Since O i N = O i M + MN , in (11.4)-(11.6) we find (after a few elementary
computations) the following expression of the absolute angular velocity:
→ n →
ω n 0= ∑ ω i ,i −1 (11.7)
i =1
The absolute angular velocity in the rigid motion with respect to fixed trihedral is
the vector sum of the relative angular velocities of the component motions.

191
Theoretical mechanics – theory and applications

11.3. Composition of accelerations

We will relate, for the beginning, the motion of the rigid to the movable
trihedrals O 2 x 2 y 2 z 2 , O 1 x 1 y 1 z 1 and to the fixed trihedral O x 0 y 0 z 0 .
The relative motion of the point M is its motion related to O 1 x 1 y 1 z 1 , hence:
→ → → → → →  →
a r= a 21+ ε 21 × O 2 M + ω
×  ω × O 2M  (11.8)
21
 21 
The transport acceleration is the acceleration related to the fixed trihedral of a
point stiff-connected to the trihedral O 1 x 1 y 1 z 1 and that shall coincide at the considered
instant with the point M, namely:
→ → → → → → → 
a t = a 10 + ε 10 × O 1 M + ω 10 ×  ω 10 × O 1 M  (11.9)
 

The Coriolis acceleration is calculated with the relative velocity v r (given by

(11.1)) and with the transport angular velocity ω 10 :
→ → → →→ → → 
a C = 2 ω 10 × v r = 2 ω
×  v + ω × O 2M  (11.10)
10
 21 21

The absolute angular acceleration of the point M shall be:
→ → → → → → → → → →
a a =
M
a r + a t + a C = a 10 + a 21 + ε 10 × O 1 M + ε 21× O 2 M +

→ → →  → → →  → → → → 
+ ω 10 ×  ω 10 × O 1 M  + ω 21×  ω 21× O 2 M  + 2 ω 10 ×  v 21+ ω 21× O 2 M  (11.11)
     

It can be demonstrated (see [2]) that, if there are more transport motions, the
absolute acceleration of the arbitrary point M is given by the relation:
→ n → n → → n → → → 
a a =
M
∑ a i,i−1 + ∑ ε i,i−1× O i M + ∑ ω i,i−1×  ω i,i−1× O i M  +
i =1 i =1 i =1
n i −1 → → → → 
+ 2 ∑ ∑ ω j , j −1×  v i ,i −1+ ω i ,i −1 × O i M  (11.12)
i =1 j =1  

and that the absolute angular acceleration ε n0 in the rigid motion related to the fixed
trihedral is:
→ n → →
ε n 0= ∑ ε i ,i −1+ ε C (11.13)
i =1
→ n i −1 → →
where ε C = ∑ ∑ω j , j −1× ω i ,i −1 is a correction term called complementary angular
i =1 j =1
acceleration. Relations (11.12) and (11.13) determine completely the distribution of
accelerations in the relative motion of the rigid. The absolute acceleration of another point

192
Theoretical mechanics – theory and applications

(point N) is obtained with the help of Euler’s formula for the rigid in general motion,
selecting point M as origin:
→ → → → → → → 
a aN = a aM + ε n 0 × MN + ω n 0 ×  ω n 0 × MN  (11.14)
 

11.4. Particular Cases. Compositions of instantaneous motions.

There are in practice particular cases of instantaneous relative motions of the rigid
stiff-connected to the trihedral O n x n y n z n with respect to other rigids to which the
trihedrals O i x i y i z i , i = 1, n − 1 were attached. We shall discuss below some of these
cases.
11.4.1. Compositions of translations
Since all of the movable trihedrals perform translation motions, we have the
following particularizations:
→ → → → →
ω i ,i −1= ε i ,i −1= 0 , v i ,i −1, a i ,i −1 = arbitrary vectors (11.15)
From the relations (11.5), (11.7), (11.12) and (11.13) it results that:
→ → → → n → → n →
ω n0= ε n0= 0 , v a =
M
∑ v i,i−1 , a a =
M
∑ a i,i−1 (11.16)
i =1 i =1
which means that the velocity and acceleration distribution correspond to a translation
motion.

11.4.2. Compositions of parallel rotations

The origins of the movable trihedrals are chosen on the corresponding axes of
rotation so that:
→ →
v i ,i −1= 0 , i = 1, n (11.17)

Denoting by u the versor of the common direction of the rotation axes, we have:
→ → → →
ω i ,i −1= ωi ,i −1 ⋅ u , ε i ,i −1= ε i ,i −1 ⋅ u , i = 1, n (11.18)
From (11.7) it follows:
→ n →  n  → →
ω n 0 = ∑ ω i ,i −1⋅ u =  ∑ ω i ,i −1  ⋅ u = ω n 0 u (11.19)
i =1  i =1 
n
where ω n 0 = ∑ ω i ,i −1 , and from (11.13) it results:
i =1
→ n →  n  → →
n 0= ∑ε i ,i −1 u = ∑ ε
ε ⋅  i ,i −1
⋅ u = ε n0 u (11.20)
i =1  i =1 
→ → n
as ε C = 0 (vector products are null). We used the notation ε n0 = ∑ε i ,i −1 .
i =1

193
Theoretical mechanics – theory and applications

The velocity of the arbitrary point M has the particular expression:


→ n → → → n →
v a =
M
∑ ω i,i−1⋅ u × O i M = u × ∑ ω i,i−1⋅ O i M (11.21)
i =1 i =1
→ → → →
which shows that v M
a ⊥ u (or v M
a ⊥ω n 0 ). Two situations are possible:
→ →
i) ω n o ≠ 0 : The distribution of velocities is that matching to a motion of rotation

at the angular velocity ω n 0 about an axis that contains the center of the parallel vectors

ω i ,i −1, i = 1, n , called instantaneous axis of rotation. Indeed, the absolute velocity of the
point M (given by (11.21)) may be written as:
n →

→  n  →
∑ ω i,i−1⋅ O i M → →
v M 
a = ∑ ω i,i−1  ⋅ u × i=1 = ω n0 × r C (11.22)
 i =1  n
∑ ω i,i−1
i =1
→ →
where r C is the position vector of the center of the parallel vectors ω i ,i −1, i = 1, n , defined
identically as the center of the parallel forces.
→ →
As ω n 0 // ε n0 (see (11.20)), the distribution of accelerations is identical to that
in the motion of rotation about the instantaneous axis of rotation.
→ →
ii) ω n o = 0 : The absolute velocities of the points M and N (relations (11.4) and
(11.5)) become:
→ n → → → n → →
v a =
M
∑ ω i,i−1 × O i M , v a =
N
∑ ω i,i−1 × O i N
i =1 i =1
Subtracting member by member the two relations it follows that:
→ → n → → → n →  → →
v aM − v aN = ∑ ω i ,i −1 × (O i M − O i N ) = ∑ ω i ,i −1  × MN = 0 (11.23)
i =1  i =1 
Relation (11.23) shows that, at an instant, the distribution of velocities is the same
as in a translation motion, all the points having the same velocity.

11.4.3. Compositions of concurrent rotations

Origins of the movable trihedrals are chosen on the axes of rotation (hence the
relation (11.17)) is verified, and coincide:
→ → → not →
O ≡ O 1≡ O 2 ≡ ... ≡ O n ⇒ O 1 M = O 2 M = ... = O n M = r
From the relation (11.4) it is deduced that:
→ n → → n →  → → →
v aM = ∑ ω i ,i −1 × O i M =  ∑ ω i ,i −1  × r = ω n 0 × r (11.24)
i =1  i =1 

194
Theoretical mechanics – theory and applications

It results that in the case of the concurrent rotations the distribution of velocities is

obtained as in a motion of rotation at the angular velocity ω n 0 about an axis passing
through O, called instantaneous axis of rotation.
Also, the distribution of accelerations is the one in the motion of a rigid with fixed
→ → → →
point, as the vectors ω n 0 and ε n0 are no longer parallel. Even if ε i ,i −1= 0 (uniform
rotations), the absolute angular acceleration is non-zero:
→ → n i −1 → →
ε n0 = ε C = ∑ ∑ω j , j −1× ω i ,i −1
i =1 j =1

and has a line of action differing from the one of the absolute angular velocity ω n 0 .

11.5. Solved problems

R 11.1) A triangular wedge ABC, with the base angles equal to α , is placed between two
→ →
bodies having translation motions on a horizontal plane at the velocities v 10 and v 20
(Figure R 11.1.1). Study the absolute motion of the wedge.
3 a
B C
2 α α 1
 
  V30
V20 V10  V32
V31

c β α α b
A  
V10 V20
0
// AC // AB

Figure R 11.1.1 Figure R 11.1.2

Solution: Denote the fixed system (horizontal plane) by (0) and the moving bodies
by (1), (2) and (3). The wedge is moving with respect to the bodies (1) and (2) which, at
their turn, are moving with respect to the horizontal plane (0). Since the component
motions are translations it results that the wedge motion shall be a translation, too, at the

absolute velocity v 30 given by the relation:
→ → → → →
v 30 = v 10 + v 31 = v 20 + v 32 (1)
→ →
In relation (1) velocities v 10 and v 20 are known and the relative velocities
→ →
v 31 and v 32 , in the wedge motion related to the bodies (1) and (2), are parallel to AC
and AB, respectively. The geometrical representation of the vector relation (1) is given in
Figure R 11.1.2.
195
Theoretical mechanics – theory and applications

From the isosceles triangle abc (similar to triangle ABC) the relative velocity
between the wedge and the bodies (1) and (2), respectively is obtained:
v −v
v31 = v32 = 20 10 (2)
2 cos α
Using the cosine theorem in the triangle abc we find the magnitude of absolute
velocity of the wedge:
1 2 + v 2 + 2v v
v30 = v 10 10 20 cos 2α (3)
2 cos α 20
The angle it makes with the horizontal plane is obtained with the sine theorem
applied in the same triangle:
v sin β v 20 − v 10
sin β = 32 = tg α (4)
v 30 2 v 30
Particular cases:
→ →
1) v 10 = v 20
From relations (3) and (4) it is obtained:
→ → →
v 30 = v 10 = v 20 , β =0 (5)
which means that the wedge and the bodies (1) and (2) are making a block moving on the
horizontal plane.
→ →
2) v 10 = - v 20
Particularizing the relations (3) and (4) we find that:
π
v 30 = v 10 = v 20 , β = (6)
2
that is the wedge shall move in the vertical direction with the same velocity the bodies (1)
and (2) approach each other.
R 11.2) Study the motion of a bearing ball in a bearing in which a truncated cone shaped
pivot is supported knowing that there is no slippage neither between the bearing and the
housing nor between the bearing ball and the shaft (Figure R 11.2.1). The angular velocity

ω 10 of the shaft and the angles α and β are known.
b

ω20
1 β−α
o1

  β
ω10 ω21
// O1D
O 2
D B α 
ω10
C
α
β a
// O1B
O1

Figure R 11.2.1 Figure R 11.2.2

196
Theoretical mechanics – theory and applications

Solution: Denote by (0) the fixed element (housing), by (1) the shaft and by (2) the
ball bearing. Since there is no slippage between the bearing and the shaft it follows that the
instantaneous axis of the relative motion bearing-shaft is the line O 1B (Figure R 11.2.1).

Let ω 21 be the angular velocity in this motion. The line of action of the angular velocity

ω 10 of the shaft contains the point O 1 as well, so that we are in the case of a composition
of two concurrent rotations. Since there is no slippage between the ball and the fixed

element we also obtain that the resultant angular velocity ω 20 , in the ball motion related
to the housing has the direction O 1D so that we may write:
→ → →
ω 20 = ω 10 + ω 21 (1)
The geometrical representation of the vector relation (1) is shown in Figure R
11.2.2. Applying the sine theorem in the triangle o 1a b it is obtained:
ω 10 ω ω 21
= 20 = (2)
sin (β − α ) sin α sin (180 − α )
hence:
sin β sin α
ω 21= ω 10 , ω 20 = ω 10 (3)
sin (β − α ) sin (β − α )

R 11.3) Study the motion of a bearing ball in a cylindrical bearing considering that the ball
rolls without slippage on the shaft journal and on bearing wall (Figure R 11.3.1). The

angular velocity ω 10 of the shaft, the journal radius R and the bearing ball radius r are
known.

ω21

ω20
2
2r

1

ω10 R 2r
2R

ω10
A

Figure R 11.3.1. Figure R 11.3.2.

Solution: Let us denote by (0) the fixed element (bearing), by (1) the shaft journal
and by (2) the bearing ball. The instantaneous axis of the ball-shaft relative motion passes

through A and is horizontal as there is no slippage between the ball and the shaft. Let ω 21

be the angular velocity in this relative motion. Since the angular velocity ω 10 of the shaft

197
Theoretical mechanics – theory and applications

related to the fixed element is horizontal, too, we have a case of composition of parallel
rotations.
In addition, since between the ball and the fixed element there is no slippage as
well, the instantaneous axis of the absolute motion of the ball passes through B. Denoting

by ω 20 the angular velocity of the absolute motion of the ball, we may write:
→ → →
ω 20 = ω 10 + ω 21 (1)
The geometrical representation of the relation (1) is shown in Figure R 11.3.2. The

resultant angular velocity ω 20 is outside the component vectors which means that the
component rotations (parallel) are of opposite senses. The resultant is closer to the
component larger in magnitude. The unknowns ω 21 and ω 20 are determined by writing
the relations from the composition of the parallel and of opposite sense vectors (that do not
form a couple):
ω 10 ω 21 ω
ω 20 = ω 21− ω 10 , = = 20 (2)
2r R + 2r R
So it is found that:
R + 2r R
ω 21= ω 10 , ω 20 = ω 10 (3)
2r 2r

R 11.4) Disc (2), of radius R 2 = R , driven by the rod OO 1 , rolls without slippage over the
disc (1) of radius R 1= 2 R (Figure R 11.4.1). The absolute angular velocities of the disc
2
(1), ω 10 = ω 0 and of the rod (3), ω 30 = ω 0 are known, where ω 0 is a positive constant.
3

Determine the absolute angular velocity ω 20 of the disc (2) and accelerations of the points
M and N of the disc.
2
 M
M aM
R2 − ω220 ⋅ O1N
 O1
• O1 aO
1

C • N
1 ω30
A ω20 N 
aO
1

aN
R1
• O
O
ω10

Figure R 11.4.1. Figure R 11.4.2.

Solution: The motions of the three bodies being rotations in a plane (bodies 1 and
3) or plane-parallel (body 2) the angular velocities are perpendicular on the plane so that
the motion of the body (2) results as a composition of parallel rotations:

198
Theoretical mechanics – theory and applications

→ → → → →
ω 20 = ω 10 + ω 21 = ω 30 + ω 23 (1)

Taking the clockwise direction for the angular velocity ω 10 we may write:
ω 20 = ω 21+ ω 10 = ω 23− ω 30 (2)
On the other hand, in the case of parallel rotations an instantaneous rotation about

the center of the parallel vectors ω results for the body (2). Denoting by C this point, from
Figure R 11.4.1. we obtain:
ω 20⋅ OC = ω 21⋅ OA = ω 23⋅ OO 1 (3)
as A is the point of null velocity between the bodies (2) and (1) and O 1 is the point of null
velocity between bodies (2) and (3). Solving the system made from the equations (2) and
(3) the values for the unknowns ω 21, ω 23, ω 20 and OC are obtained as follows:
10 5
ω 21= −5 ω 0 , ω 23= − ω 0 , ω 20 = −4 ω 0 , OC = R (4)
3 2
The accelerations of the points M and N are obtained from the relations:
→ → → → → → → 
a M = a O 1 + ε 20 × O 1M + ω 20 ×  ω 20 × O 1M 
 
 
→ → → → →  → → 
a N = a O 1 + ε 20 × O 1 N + ω 20 ×  ω 20 × O 1 N  (5)
 
 
But:
→ →
→ ∂ ω 20 ∂ ω 21 → → →
ε 20 = + + ω 10 × ω 21 = 0 (6)
∂t ∂t
→ →
as the vectors ω 10 and ω 21 are constant and parallel and:
→ → →  → → → →  →
ω 20 ×  ω 20 × O 1M  = −ω 220⋅ O 1M , ω 20 ×  ω 20 × O 1 N  = −ω 220⋅ O 1 N (7)
   
   
we obtain the following relations for determining accelerations of the points M and N :
→ → → → → →
2 ⋅O M , a
a M = a O 1 − ω 20 2
1 N = a O 1 − ω 20⋅ O 1 N (8)

Vector a O 1 has the direction OO 1 , sense from O 1 to O and the magnitude:
4
a O 1 = OO 1⋅ ω 220 = R ω 02 (9)
3
The directions and senses of the accelerations of the points M and N, determined
as per the relation (8), are shown in the Figure R 11.4.2. and their magnitudes are:

52 2320
a M = a O 1 +ω 2 2 2 4 2
20⋅ O 1M = 3 R ω 0 , a N = a O 1 + ω 20⋅ O 1 N = R ω 02 (10)
3

199
Theoretical mechanics – theory and applications

11.6. Problems proposed


TC 11.1) Determine the velocity of the rod (2) of the mechanism in the Figure TC 11.1 if

we know the velocity v 1 of the translation follower (1) and the inclination angle α
thereof.
TC 11.2) Given the planetary mechanism in the Figure TC 11.2 consisting in the gear
wheels (1), (2) and the crank (3). Wheel (2) rolls without slippage over the wheel (1) while
rolling inside a cylindrical toothed surface coaxial with the wheel (1). The radiuses R and
r, respectively, of the wheels (1) and (2), respectively and the angular velocity ω 0 = ε 0 ⋅ t
of the crank (3) are known, where ε 0 is a positive constant. Determine:
a) The absolute angular velocities and accelerations of the wheels(1) and (2) ;
b) The absolute velocity and acceleration of the point M.


V1
2
1
r A
• ⋅

 α 1
M
V21
3

 R O
V2 •  
2 ω30 = ω0

Figure TC 11.1 Figure TC 11.2



TC 11.3) A disc of radius r rotates at the constant angular velocity ω 2 about the arm CD

which, also, rotates at the constant angular velocity ω 1 about the axis AC (Figure TC
→ →
11.3). Determine the angular velocity ω and the angular acceleration ε in the motion of
→ →
the disc D as well as the velocity v B and acceleration a B of the point B located at the
disk periphery.
y y
ω1
R P
ω1
θ x

C x r
ω2

D h
A r z
ω2 B
z

Figure TC 11.3 Figure TC 11.4


200
Theoretical mechanics – theory and applications

TC 11.4) A disc of radius r fixed in a fork (Figure TC 11.4) rotates at the constant angular

velocity ω 2 around its axis while the fork rotates around its own axis at the angular
→ →
velocity ω 1 . Determine the absolute angular acceleration ε of the disc as well as the
acceleration of the point P at the disc periphery when θ = 0 0 and θ = 900 , respectively.

11.7. Indications and answers

TC 11.1) v 2 = v 1 sin α , v 23= v 1 cos α ;


2(r + R ) r+R 2(r + R ) r+R
TC 11.2) a) ω 10 = ω 0 , ω 20 = ω 0 , ε 10 = ε 0 , ε 20 = ε 0;
r r r r
R+r
b) v M = 2 (R + r ) ε 0 t , a M = ε 0 ε 02 t 4 r 2 + (R + r )2  + r 2 1 − 2 ε 0 t 2  ;
r    
→ → → → →
TC 11.3) ω = ω 1 j + ω 2 k , ε = ω 1ω 2 i ;
→ → → → →
v B = r ω 2 j − (R + r )ω 1 k , a B = − (R + r )ω 12 + r ω 22  i ;
 
→ →
TC 11.4) ε = −ω 1 ω 2 j ;
→ → →
a P == − r ω 22 i + 2 r ω 1 ω 2 k , when θ = 0 0 ;
→ →
a P = − r  ω 12 + ω 2  j , when θ = 900 .
 2

201
Theoretical mechanics – theory and applications

DYNAMICS
Dynamics is concerned with the study of the motion of material points or rigid
solid systems taking into account the mass and the forces acting on the analyzed system
components.
General problems of dynamics are two in number, namely:
i) Knowing the forces acting on the material system and the initial conditions (system
configuration and velocity distribution at the initial instant) it is requested to determine the
motion of the system;
ii) Knowing the motion of the system, determine the forces acting on it.

We will begin the study of dynamics by analyzing the motion (absolute and relative) of
the material point (free and constrained).

12. Dynamics of the absolute motion of the material point


12.1. Dynamics of the free material point

The general problems of free material point dynamics can be analyzed using the
second principle of mechanics, written as:
→ →
ma=F (12.1)

where m is the mass of the material point (constant during motion), a its acceleration and
→ →
F the resultant of the forces acting on the point. In general, the force F is a function of the

→ → →
position vector r , velocity v = r and time t, that is:
 • 
→ → → → 
F = F r , r , t  (12.2)
 
 
••
→ →
Remarking that a = r , the fundamental equation (12.1) is written as:
••  • 
→ → → → 
m r = F r , r , t  (12.3)
 
 
i) Direct problem: The forces acting on the material point are known, its motion
is requested.
To study the motion, the vector equation (12.3) is integrated. In solving concrete
problems of mechanics, the equation (12.3) is projected on the axes of a coordinate
system of convenient choice. Thus we have:

202
Theoretical mechanics – theory and applications

Cartesian coordinate system:


••  • • •  ••  • • •  ••  • • • 
m x = F x x, y, z , x, y, z , t  , m y = F y  x, y, z , x, y, z , t  , m z = F z  x, y, z , x, y, z , t  (12.4)
     
Frenet coordinate system:

••  •  s2  •   • 
m s = F τ  s, s, t  , m = F ν  s, s, t  , 0 = F β  s, s, t  (12.5)
  ρ    
Polar coordinate system:
•• •
2  • •   •• • •  • • 
m r − r θ  = F r  r , θ , r , θ , t  , m r θ + 2 r θ  = F n  r , θ , r , θ , t  (12.6)
       
The basic equation (12.3) can be projected as well on the axes of other reference
systems (spherical, cylindrical, generalized, and so on). The different forms of differential
equation thus obtained represent a system of three second order differential equations with
three scalar functions of time t as unknowns (with the Cartesian coordinates, they are x(t),
y(t) and z(t)). Considering that the system can be integrated, the resulting solutions will
depend on the time t and on six constants of integration. We will use the most commonly
used coordinate system, namely the Cartesian coordinates one. The solution of the system
(12.4), called the general solution is of the form:
x = x(t , C 1, C 2 , C 3 , C 4 , C 5 , C 6 ) ,
y = y (t , C 1, C 2 , C 3 , C 4 , C 5 , C 6 ) , (12.7)
z = z (t , C 1, C 2 , C 3 , C 4 , C 5 , C 6 ) .
In applications, of interest is the solution called particular solution that satisfies
the initial conditions of motion concerning the position and velocity of the point at the
initial instant.
• • •
t = 0 : x = x 0, y = y 0, z = z 0, x = v 0x , y = v 0 y , z = v 0z (12.8)

Imposing the conditions (12.8), from (12.7) results:


x 0 = x(0, C 1, C 2 , C 3 , C 4 , C 5 , C 6 )
y 0 = y (t , C 1, C 2 , C 3, C 4 , C 5, C 6 )
z 0 = z (t , C 1, C 2 , C 3, C 4 , C 5, C 6 )

v 0 x = x (t , C 1, C 2 , C 3 , C 4 , C 5 , C 6 )

(12.9)

v 0 y = y (t , C 1, C 2 , C 3 , C 4 , C 5 , C 6 )

v 0 y = z (t , C 1, C 2 , C 3 , C 4 , C 5 , C 6 )

The six equations (12.9) constitute a system of algebraic equations in the


unknowns C i , i = 1,6 . Solving this system we obtain the values of the integration constants
depending on the initial data of the motion:
( )
C i = C i x 0 , y 0 , z 0 , v 0 x , v 0 y , v o z , i = 1,6 (12.10)
Substituting (12.10) in (12.7) we obtain the particular solution:

203
Theoretical mechanics – theory and applications

(
x = x t , x0 , y 0 , z 0 , v 0 x , v 0 y , v 0 z ) ,
(
y = y t , x0 , y 0 , z 0 , v 0 x , v 0 y , v 0 z ), (12.11)
z = z (t , x , y
0 0, z 0, v 0 x , v 0 y , v 0 z ).
The set of relations (12.11) correspond to the parametric equations of the motion
(see chapter “Kinematics of the relative motion of material point”).

ii) Reversed problem: The point motion is known; the forces acting on it are
requested.
• ••
→ → → → → →
Given the vector r = r ( t ) we obtain the vector functions v = r ( t ) and a = r ( t ) .
→ →
Equation (12.3) allows sole determination of the resultant F = m a , but this expression
does not give us any indication on the physical nature of the resultant (there is infinity of
force systems having the same resultant).

12.2. Dynamics of the motion of the constrained material point

In the study of motion of a constrained material point the procedure is as in statics,


more precisely the connections axiom is applied substituting each constraint with the
matching reaction and treat the problem as if the material point were free. Denoting by
→ →
R ' the resultant of the connection forces and by F the resultant of the directly applied
forces, the equation of motion of the constrained material point is:

→ → →
m a = F+ R ' (12.12)

By projecting this equation on the axes of a reference system conveniently chosen


we obtain a system of second order differential equations whose unknowns concern the
motion and the connection forces. To determine the motion, the variation in time of a
position parameter (for point on the curve) or two position parameters (for point on the
surface) is determined. Determination of the reactions is usually done by projecting
equation (12.12) on the normal or tangent direction to the support curve or surface.

12. 3. Solved problems

R 12.1) Determine the initial velocity v 0 and angle for the projectile to be launched from
point O to hit normally at height AB =h a vertical wall AB located at the distance OA = 2h.
Neglect air opposition. (Figure R 12.1.1).

204
Theoretical mechanics – theory and applications

y
B

VB
B ( 2h , h )
 M
V0 •
h  
V0 mg
α
O A α x
O
A
2h

Figure R 12.1.1 Figure R 12.1.2

Solution: Let x and y be the coordinates of the point M at instant t (Figure R


12.1.2). We have:
 ••  x(t ) = C 1 t + C 2
→ → m x = 0 
m a = G ⇒  •• ⇒  gt2 .
m y = −m g  y (t ) = − +C3t+C4
  2
Using the initial conditions:
 x = 0, y = 0
t=0 :  • •
v x = x = v 0 cos α , v y = y = v 0 sin α
the constants are obtained as follows:
C 1= v 0 cos α , C 2 = 0, C 3 = v 0 sin α , C 4 = 0 .

The motion is described by the parametric equations as:


gt2
x( t ) = v 0 t cos α , y (t ) = − + v 0 t sin α .
2
By eliminating the time t the motion curve is obtained (a parabola):
g
(C ) : y ( x) = − x 2 + x tg α .
2 v 0 cos α
2 2

But B( 2h, h) ∈ (C ) leads to y ( 2h) = h,


dy
( 2h) = 0 (*). The second condition (*)
dx
is the consequence of the horizontality of the velocity vector (tangent to the trajectory) at
the point of impact with the wall. Conditions (*) lead to the system:
g g
− ( 2h) 2 + 2h tg α = h , − 2 ( 2h) + tg α = 0 ,
2 v 0 cos α
2 2
v 0 cos 2 α
with the solution:
π
v 0= 2 g h , α= .
4

R 12.2) Point M, of mass m, moves frictionless on an incline at the angle α with the
horizontal and is attracted to the point O with a force proportional to the distance OM, the
coefficient of proportionality being K = k ⋅ m , where k is a positive constant (Figure R

205
Theoretical mechanics – theory and applications

12.2.1). Knowing that at the initial instant the mobile M is at rest at A and the distance AO
= l, it is requested:
a) The law of motion AM = x(t) and the value of the reaction normal to the plane;
b) The mobile M velocity at the point B.

 A 

A N i j
t)
x(
M
 • 
l y
a F
l 
G
M
B α O
α
B O x

Figure R 12.2.1 Figure R 12.2.2

Solution: a) The position of the mobile at the instant t is given by AM = x(t). The
fundamental equation of the dynamics is written as:
→ → → →
m a = G+ F + N (1)
→ → →  → →
where N is the normal reaction and F = − K ⋅ OM = − k m  ( x − l sin α ) i − l cos α j  is
 
the force the point M attracts the mobile with (Figure R 12.2.2). Projecting equation (1) on
the axes of the reference Axy it follows:
 ••
 m x = m g sin α − k m (x − l sin α ) (2)
0 = m g cos α + k m l cos α − N

The initial conditions of the motion are:



t = 0 : x = 0 ,v = x = 0 (3)
From the first equation (2) and from (3) the law of motion results:
g+kl
x(t ) = AM =
k
(
1 − cos k t sin α ) (4)
and from the second equation (2) the normal reaction on the plane:
N = m (g + k l ) cos α (5)

b) The time required to travel the distance AB is given by the relation:


g+kl
AB =
l
sin α
= x (t B ) =
k
(
1 − cos k t B sin α ) (6)
It follows:
kl
cos k t B = 1 −
(g + k l) sin 2 α

206
Theoretical mechanics – theory and applications

• ( 4) g+kl
v(t ) = x ( t ) = sin k t sin α (7)
k
g+kl
2 l (g + k l ) sin 2 α − k l
1
v B = v(t B ) = sin k t B sin α = 2
.
k sin α

R 12.3) A pointlike ball of mass m is tied through an inextensible cord of length l and
moves in a horizontal plane with the coefficient of friction µ (Figure R 12.3.1).
a) Determine the value of tension in the cord and of the velocity of the ball in an
arbitrary position given by the angle θ if the ball initial launching speed is perpendicular
to the direction of the stretched cord and has the value v 0 ;
b) What should be the value of the speed v 0 if after completing a circumference the
ball stops?
β

ν 

O N τ
O S
l 
l θ
θ V
M0 ⋅ M0
•M
 M ( m) 
V0 T

mg
Figure R 12.3.1 Figure R 12.3.2

Solution: a) The projections of the fundamental equation on the axes of the Frenet
trihedral Mτ ν β (Figure R 12.3.2) are:
••
Mτ : m s =−T (1a)

2
s
Mν : m =S (1b)
l
Mβ : 0 = N − m g (1c)
T=µN (2)
where s = l θ is the length of the arc M 0 M , N is the normal reaction, T is the slippage
friction force and S is the tension in the cord.
From (1a), (1c) and (2), by integration it results:
µ g 2
s (t ) = − t + C1 t + C 2 (3)
2
The initial conditions are:

t =0 : s=0 , s=v0 (4)

207
Theoretical mechanics – theory and applications

From (3) and (4) we find the values C 1= v 0 , C 2 = 0 , thus the laws of motion on
the circle C(O, 1) are:
µg 2 •
s=v0 t − t , v = s = v 0− µ g t , a τ = − µ g (5)
2
From the first two equations (5) we obtain the velocity in a certain position given
by the angle θ :
v 2 = v 02 − 2 µ g s = v 02 − 2 µ g l θ (6)
and from (1b) the expression of the tension in the cord:
S (θ ) =
m 2
l 0
(
v −2 µ glθ ) (7)
b) The condition for motion cease after completing a circumference:
θ = 2 π ⇒ v (θ ) = 0 (8)
gives the value of the launching speed:
v 0= 2 π µ g l (9)

R 12.4) A material point of mass m is attracted by the attractive center O by a force


→ 5m →
inversely proportional to the cube of the distance from it to the center O, F = − 4 r . At
r
the initial time the mobile is at point M 0 given by OM 0 = r 0 and has the value v 0 ,
inclined at the angle α with the direction OM 0 (Figure R 12.4). Requested:
a) The differential equations of the motion in polar coordinates;
b) Areal constant;
c) Parametric equations of the motion;
d) Velocity components in polar coordinates;
e) Trajectory equation in polar coordinates.

ρ

F •M

r 
V0

n
α0
θ
O r0 M0

Figure R 12.4

→ →
Solution: a) The fundamental equation m a = F is projected on the axes of the
polar coordinates system ρ O n :
•• •
2 5m  •• • •
m r − r θ = − 3 , m r θ + 2 r θ  = 0 (1)
  r  

208
Theoretical mechanics – theory and applications

b) Multiplying the first differential equation by r we obtain:


d  2• •
 r θ  = 0 ⇒ r 2 θ = r v n = C = constant (2)
dt  
Constant C is called areal constant and is obtained from the initial conditions:
t = 0 : v n = v 0 sin α = 1 m / s , r = r 0 = 2 m (3)
Results that: C = r 0 v 0 sin α = 2 m 2 / s .
•• 1
c) From the first equation (1) and from (2) the differential equation r + 3
= 0 is obtained,
r

which is multiplied by 2 r to determine the unknown r(t):
d  • 2 1  • 1 • C 1r 2 + 1 16 − t 2
− = ⇒ 2
− = ⇒ = ± ⇒ ( ) =
dt  
r 0 r C 1 r r t (4)
r 2 r2 r 2
Variation in time of the polar angle θ is obtained from the initial conditions (3)
and equation (2). It is found that:
4+t
θ ( t ) = ln (5)
4−t
• t • 4
d) v ρ = r = − , v n= r θ = (6)
2 16 − t 2 16 − t 2
e) Eliminating the time t between the parametric equations (4) and (5) the trajectory
equation in polar coordinates is obtained:
θ 
4 exp 
2
r (θ ) = (7)
exp θ ) + 1
(
which is a spiral in the plane ρ O n .

12.4. Problems proposed

12.4.1. Conventional tests

TC 12.1) A material point (mobile) P, of mass m, moves in the vertical plane Oxy, under
→ →
the action of its own weight and the elastic force F = − k ⋅ OP , where k is a positive
constant (Figure TC 12.1.1). At the initial instant the mobile is at point A on the vertical

axis (OA = h) and has the velocity v 0 horizontal. Given the values m, k, h, v 0 , determine:

a) Differential equations of motion;


b) Parametric equations of motion;
c) Motion trajectory;
π m
d) Position, velocity and acceleration of the mobile at the instant t 1= .
2 k

209
Theoretical mechanics – theory and applications

y

A V0 m
P ( x , y) 
 µ V0
F  
G = mg
k
h

j
α

O i x

Figure TC 12.1.1 Figure TC 12.2.1

TC 12.2) On a plane inclined at the angle α with the horizontal line there is a body of
mass m and a spring of elastic constant k (Figure TC 12.2.1). Between the body and the
plane there is a coefficient of friction µ . At the initial instant when the body is in contact
with the spring in natural condition (neither tensed nor compressed) the body is given the

velocity v 0 in the direction of spring compression. Requested:
a) Law of motion for the descending body;
b) Motion duration (till body stops);
c) Distance traveled till stop.
It is supposed that µ > tg α .

TC 12.3) A material point of mass m is moving frictionless on a fixed cylinder of radius R


(Figure TC 12.3). At the beginning of the motion the mobile was at point A and had the

initial horizontal velocity v 0 . Requested:
a) Determine the laws of variation v = v ( θ ) , a τ = a τ ( θ ) , a ν = a ν ( θ ) for the
time interval the mobile stays in contact with the cylinder;
not
b) Angle θ = α at which the mobile detaches from the cylinder surface;
not
c) For the particular case v 0 = 0 , determine the distance OC = ξ , point C is the
point where the mobile hits the ground.
A ( t = 0)  
• V0 N

• M ( t > 0)
y


α G •D
θ 
O′ VD

D′ C x
O

Figure TC 12.3

210
Theoretical mechanics – theory and applications

12.4.2. Multiple-choice tests

TG 12.1) A crane is lifting upright a car weighing G = 10 kN for the time t = 10 s , at the
height h = 12 m . It is assumed that the initial velocity of the weight is null and
that g = 10 m / s 2 . The force intensity, assumed as constant, applied to the load is:
a) F = 10,24 kN ; b) F = 12 kN ; c) F = 9,5 kN ; d) F = 11,95 kN .

TG 12.2) Intensity of the force F, required to decelerate an automobile of 7 kN in order


that after 10 s its speed of 60 km/h is reduced to a half is:
a) 325 N; b) 175 N; c) 594, 6 N; d) 372 N.

12.5. Indications and answers

TC 12.1) a) The expression of the elastic force depending on the coordinates x and y of the
→ → → → → →
point P is F = − k x i − k y j . The fundamental equation m a = F + G is projected on Ox
and Oy and a second order linear differential equations with constant coefficients are
obtained as follows:
•• ••
x + p2 x = 0 , y+ p 2 y = − g (1)
not
k
where p 2
= .
m
a) Integrating the system (1) into the initial conditions :
• •
t =0 : x=0, y = h, x=v0 , y =0 (2)
we find the parametric equations:
v0  g  g
x( t ) = sin p t , y ( t ) =  h + 2  cos pt − 2 (3)
p  p  p
b) The trajectory is the ellipse of equation:
 g 2
 y + 
x2  p2 
+ =1.
 v 0 2  g 2
   h + 
 p 2
 p 
Its representation is shown in Figure TC 12.1.2.
c) See “Kinematic of the absolute motion of the material point”.
v0 g
Position: x = , y=− 2 .
p p
 g 
Velocity: v x = 0 , v y = − p  h + 2  .
 p 
Acceleration: a x = − v 0 p , a y = 0.

211
Theoretical mechanics – theory and applications

y O
A (0, h ) x

Fe
y 
 N
T
 
a1 V0 a
D( ,−
g
) α
p p2 
C(0,−g / p 2 ) 
mg
v1

Figure TC 12.1.2 Figure TC 12.2.2

TC 12.2) a) The projections of the fundamental equation on the axes of the reference Oxy
(Figure TC 12.2.2) are:
••
m x = m g sin α − T − Fel , 0 = N − m g cos α (1)
where N is the normal reaction, T = µ N friction force and Fel = k x the elastic force
corresponding to the spring deformation. From (1) the second order inhomogeneous
differential equation is obtained:
•• k
x+ x = g (sin α − µ cos α ) (2)
m
with the solution:
 
x( t ) = v 0
m
sin
k
t+
gm
(sin α − µ cos α )1 − cos k t  . (3)
k m k  m 

b) Let t = t 1 be the motion duration. The condition: t = t 1, v = x = 0 is imposed. It results:
m v 0 k 1 
t 1= arctg  (4)
k  g m µ cos α − sin α 
c) D = x (t 1 ) , where x (t 1 ) is determined from (3) and (4).

 ∧ 
TC 12.3) a) At an arbitrary instant t the mobile is at the point M, so that m AOM  = θ .
 
The work-kinetic energy theorem, applied between the points A and M allows obtaining
the dependence of the mobile velocity on the angle θ :
v = v 02 + 2 g R (1 − cos θ ) (1)
→ → →
The fundamental equation m a = G + N , is projected on the tangent to the
trajectory and the variation of the tangential velocity with the angle θ is obtained:
a τ = g sin θ (2)
The normal component of the acceleration results from the kinematic
v2
equation a ν = . Therefore:
R

212
Theoretical mechanics – theory and applications

v 02
aν = + 2 g (1 − cos θ ) (3)
R
b) Dependence of the normal reaction on angle θ : is obtained by projecting the
fundamental equation on the normal to the cylinder at M. It results:
m v 02
N = m g ( 3 cos θ − 2) − (4)
R
At the instant of detachment, the normal reaction is annulled (the mobile does not
push on the cylinder) and θ = α , so that: from (4) we find that:
v 2+ 2g R
α = arccos 0 (5)
3g R
2g R
c) Assuming v 0 = 0 , the velocity of the mobile at the point of detachment D is v D =
3
and the angle made by the velocity vector with the vertical direction
2
is 90 − α = 90 − arccos . As of the instant of detachment (considered as instant t = 0) till
3
hitting the ground at C, the sole force acting on the mobile is its own weight. The
→ →
fundamental equation m a = G and the initial conditions:
t = 0 : x = 0 , y = R(1 + cos α ) , v x = v D cos α , v y = −v D sin α ,
allow determination of the distance D’C till hitting the ground:
(
45 2− 5 )
D 'C = R (6)
27
5
As OD ' = R sin α = R , we find that:
3
(
5 4 2+ 5 )
ξ = OC = OD '+ D ' C = R (7)
27
TG 12.1) The motion of the weight is uniformly varied rectilinear at the
F −G at2
acceleration a = . Using the law of space in this motion, we deduce that h = ,
m 2
hence:
 2 h 
F = G 1 + 2
= 10,24 kN .
 gt 
Correct answer: a).
TG 12.2) The motion of the automobile is uniformly decelerated rectilinear at the
F F
acceleration a = − = − g . Adding the law of velocity in this motion v = v 0 + a t , we
m G
obtain:
F=
G
(v 0 − v) = 7000 ⋅ ( 60 − 30) ⋅ 1000 = 594,6 N .
gt 9,8 ⋅ 10 3600
Correct answer: c).

213
Theoretical mechanics – theory and applications

13. Dynamics of the relative motion of the material point


13.1. Fundamental equation of the relative motion

The fundamental equation of the absolute motion of the material point


→ →
ma=F (13.1)

describes its motion relative to a fixed reference. In the relation (13.1), a is the absolute
acceleration. From the kinematics of the relative motion of the material point it is known
that:
→ → → →
a a= a r + a t + a C (13.2)

→ ∂vr → → → → → → →
where a r = is the relative acceleration, a t = a O + ε × r + ω ×  ω × r  is the
∂t  
→ → →
transport acceleration and a C = 2 ω × v r is the Coriolis acceleration. From (13.1) and
(13.2) it is obtained:
→ → →  → → →  →   → 
m  a r + a t + a C  = F ⇔ m a r = F + − m a t  + − m a C  (13.3)
     
Denoting:
→ → → →
−m at = Ft , − m aC = FC (13.4)
from (13.3) we find:
→ → → →
m a r= F+ F t+ F C (13.5)
Equation (13.5) represents the fundamental equation of the dynamics of the
→ →
relative motion of the material point. Vectors F t and F C , having the dimension of forces
are called complementary transport force and complementary Coriolis force. They correct
→ →
the fundamental equation m a = F when the study of the motion is made with respect to a
mobile reference.

13.2. Relative rest

It is possible for the material point to be in equilibrium (relative) related to the


movable reference and to move jointly with it with respect to the fixed one. Therefore it is

required that its relative motion v r is null, hence the consequence:

→ ∂vr → → →  → →  →
ar = =0 , F C = −m a C = −m 2 ω × v r  = 0 (13.6)
∂t  

214
Theoretical mechanics – theory and applications

From (13.5) and (13.6) the relative rest condition is obtained:


→ → →
F + F t= 0 (13.7)
that is the material point remains at rest related to the movable reference in that position
wherein the directly applied forces and the complementary transport force make their
equilibrium.

13.3. Inertial systems

We shall deduce in this section the conditions for the fundamental equation of the
dynamics to have the same form ((12.1)) irrespective of the fact it is written with respect to
the fixed or the movable reference.
Therefore the following condition is imposed:
→ → →
F t+ F C = 0 (13.8)
or
→ → → → → → → → →
a O + ε × r + ω × ω × r  + 2 ω × v r = 0 (13.9)
 
Condition (13.9) shall have to be verified irrespective of the point wherein the
movable item finds and of its relative velocity. Considering two material points having the
→ → →
same position (same r ) but different relative motions ( v 'r and v 'r' ) and imposing them
the condition (13.9) we find:
→ → → → → → → → →
a O + ε × r + ω ×  ω × r  + 2 ω × v 'r = 0
 
→ → → → → → → → →
a O + ε × r + ω ×  ω × r  + 2 ω × v 'r' = 0 .
 
→ → →  → → →
Subtracting these relations we obtain 2 ω ×  v 'r − v 'r'  = 0 and since v 'r ≠ v 'r' it
 
results:
→ →
ω=0 (13.10)
In addition,

dω →

ε = =0 (13.11)
dt
Conditions (13.10)-(13.11) make the relation (13.9) to be reduced to:
→ →
a O= 0 (13.12)
As a conclusion, for the fundamental equation of the dynamics to keep its form
when it is also written with respect to a movable reference it is required that the movable
→ → → → →
system is in a uniform rectilinear( a O = 0 ) translation motion ( ω = ε = 0 ) related to the
fixed reference, or in particular, is at rest.
Reference systems satisfying this condition are called inertial systems.

215
Theoretical mechanics – theory and applications

13.4. Solved problems

R 13.1) The OAB angled bar rotates with the constant angular velocity ω 0 in the
horizontal plane. A ring M of mass m moves without friction on the portion AB (Figure R
 ∧ 
13.1.1). Knowing that mOAB  = 90 0 , AB = l and that at the initial instant the ring was
 

released at the velocity v 0 from point A, determine:
a) The law of motion of the ring on the bar AB;
b) The value of the reaction of the bar for an arbitrary position of the ring on the bar.
y
y
O
ω0 
O FC

Ny x
ω0 α 
l x
x
⊗ G
B
B
• z 
M
M • Ft
⋅  A Nz
z• V0
A

Figure R 13.1.1 Figure R 13.1.2

Solution: a) The relative motion of the ring M is a rectilinear motion along the bar
• ••
AB, characterized by the relative speed v r = x and relative acceleration a r = x . The
transport motion is a uniform circular motion at the angular speed ω 0 , on a circle of center
O and radius OM = l 2 + x 2 . The transport acceleration has only the normal component
of magnitude a t = OM ⋅ ω 02 (Figure R 13.1.2).
Releasing the ring M from its connection to the bar we obtain the following vector
equation of the relative motion:
→ → → → → →
m a r = G+ N y + N z + F t + F C (1)
where:
→ →
G = − m g k - is the weight force;
→ → → →
N y= N y j , N z = N z k - are the normal projections in the directions Ay and Az;
→ → → →
F t == m a t = m a t sin α i − m a t cos α j - is the complementary transport force;
→ → → → →
F C = − m a C = −2 m ω × v r = 2 m ω 0 v r j - is the complementary Coriolis force.

Projecting the vector equation (1) on the axes of the trihedral Axyz we obtain the
following scalar equations:

216
Theoretical mechanics – theory and applications

••
m x = m a t sin α
0 = N y − m at cos α + 2 m ω 0 v r (2)
0 = −m g + N z
x l
Remarking that sin α = , cos α = , from the first equation of the
l 2+ x 2 l 2+ x 2
system (2) we find the differential equation of the ring motion on the bar:
••
x − ω 02 ⋅ x = 0 (3)
whose general solution is:
x( t ) = A ⋅ exp(ω 0 t ) + B ⋅ exp(− ω 0 t ) (4)

The integration constants result from the initial conditions t = 0 : x = 0, x = v 0 . We
v0
find the values A = − B = , so that the law of relative motion of the ring on the bar is:
2ω 0

sh (ω 0 t )
v0
x( t ) = (5)
ω0
b) From the last two equations of the system (2) we obtain the normal reaction projections:
(
N y = m l ω 02 − 2 ω 0 v 0 ch(ω 0 t ) ) , N z= m g (6)
so that the value of the normal reaction at the instant t is given by the relation:
[
N = N 2y + N 2z = m g 2 + l ω 02 − 2 ω 0 v 0 ch(ω 0 t ) ]2 (7)

R 13.2) A horizontal disk rotates with the angular velocity ω 0 around its axis. At a given
instant a material point of mass m is placed freely on the disc at the radius r 0 from the
axis and having zero initial velocity relative to the disc surface (Figure R 13.2.1).
Neglecting the friction between the point and the disk, determine the motion of the point
relative to the disk surface.

Vrel 
Ft


 α Vr
M θ Vθ

r FC

O
A 
N • ⊗

r G
θ
O• A
ω0

 r0
ω

Figure R 13.2.1 Figure R 13.2.2

217
Theoretical mechanics – theory and applications

Solution: The study of the material point motion shall be made in a polar
coordinate system ( r , θ ) , in which r = OM is the distance from the point to the axis
 ∧ 
and θ = mas AOM  is the angle between the “radius” of the point M with the initial
 
direction OA (Figure R 13.2.2). The vector equation of the relative motion is:
→ → → → →
m a r = G+ N + F t + F C (1)
Since the angular velocity of the disk is constant it follows that the transport
acceleration shall have only the normal component a tν = r ω 02 non zero, so that the
complementary transport force will be directed along the radius OM, from O to M and
shall have the magnitude F t = m r ω 02 . The Coriolis complementary force is given by the
relation:
→ → → → → → 
F C = −2 ω × v relativa = −2 m ω ×  v r + v θ  (2)
 
→ →
where v r and v θ are the relative velocity components on the polar coordinate system
axes. The complementary force is, therefore, perpendicular to the relative velocity vector,
is in the disk plane, and has the magnitude:
•  • 
F C = 2 m ω 0 r  2 + r θ  2 (3)
   
Projecting equation (1) on the polar coordinate system axes we obtain the
equations:
•• •   •• • •
m  r − r θ 2  = F t − F C sin α , m  r θ + 2 r θ  = F C cos α (4)
   
Remarking that:
• •
rθ r
sin α = , cos α = (5)
• 2  •  2 • 2  •  2
r  + r θ  r  + r θ 
       
the equations (4) can be brought to the form:
•• • •• • •
r − r θ 2 = r ω 02 − 2 r ω 0 θ , r θ + 2 r θ = 2 ω 0 (6)
Multiplying by r the second equation of the system (6) it will be rewritten as:
d  2 • d 2
d t
r θ  =
 dt
(
r ω0 ) (7)

hence, by integration, we find:


• C
θ =ω 0 + 2 (8)
r
The initial conditions:
 r = r 0 , θ = 0
t = 0 : • • (9)
 r = 0 , θ = 0

218
Theoretical mechanics – theory and applications

allow us to obtain the value of the constant C, namely C = −ω 0 r 02 , hence the equation:
 r 02 

θ = ω 01 − 2  (10)
 r 
From (10) and the first equation of the system (6) we deduce the following
differential equation with the unknown r(t):
•• ω 2 r 4
0 0
r− =0 (11)
r3

Multiplying the latter by 2 r , it becomes:
d  • 2 ω 02 r 04 
d t 
r + = 0 (12)
r2 
hence, after finding the integration constant, we find that:
• r 02
r = ω 0r 0 1 − (13)
r
Integrating the severable variable equation (13) we find the dependence r = r(t):
r = r 0 1 + ω 02 t 2 (14)
Eventually, from the relations (10) and (14) we obtain the second parametric
equation of the material point motion in polar coordinates:
θ ( t ) = ω 0 t − arctg (ω 0 t ) (15)

R 13.3) A conical vessel with the angle at vertex 2 α rotates about its axis of symmetry at
the constant angular velocity ω . A ball of mass m is resting on the internal wall of the
vessel (Figure R 13.3.1). Determine the positions of relative equilibrium of the ball and the
normal reaction at these points.

x
x
ω
ω

N 
M Ft
r •A •
• M ( m)

G
2α 2α x
y z y

O

Figure R 13.3.1 Figure R 13.3.2

Solution: Releasing the ball from its connection to the vessel wall and introducing

the normal reaction N , the equation of relative equilibrium has the form:
→ → → →
G+ N + F t = 0 (1)

219
Theoretical mechanics – theory and applications

We attach the movable reference system Oxyz to the moving conical vessel (Figure
R 13.3.2) having the axis Ox along the slant on which the ball finds at relative rest at the
distance OM= x. The transport motion is a circular motion at the constant angular speed ω ,
on the circle of center A and radius r = x sin α and, therefore, the transport acceleration
has only the normal component a tν = r ω 2
non zero. The complementary transport force

Ft shall be horizontal, opposite in sense to the transport acceleration and of
magnitude F t = m r ω 2 .
Projecting the vector equation (1) on the axes Ox and Oy the following scalar
equations are obtained:
m x ω 2 sin α − m g cos α = 0 (2)
N − m g sin α = m r ω cos α = 0 2
(3)
g tg α
From equation (2) it is obtained x = = constant, meaning that the ball is in
ω2

relative equilibrium at all of the points of the circle of radius r = x sin α , where the
distance x is given by the equation above. From equation (3) we obtain the value of the
normal reaction at the relative equilibrium points:
mg
N= (4)
sin α
R 13.4) A material point M, found in the northern hemisphere, is allowed to fall freely
from the height H that is small compared to the radius of the Earth. Considering that the
weight P = mg of the point is constant during the fall, and neglecting the air resistance,
determine the deviation ε to the east, when the point touches the ground, with respect to
the vertical passing through the initial position M 0 (Figure R 13.4).
N y
y M0

•M 0
 
O
• x
ω v ε

H  λ


M F C


P
O ε x (est )

Figure R 13.4

Solution: A Cartesian reference system is chosen with the axis Oy oriented


towards the vertical direction of the initial position of the material point (towards the
radius of the Earth) and the axis Ox oriented towards east.

220
Theoretical mechanics – theory and applications

To take into account the rotation of the earth as well, besides the weight P, we
shall also enter the complementary transport and Coriolis forces. In calculation of the
complementary Coriolis force we shall neglect the component v x of the relative velocity of
the material point, since it is much smaller than the component v y . Thus, v r = v y (towards
the center of the Earth, C). In addition, we shall also consider as negligible the
complementary transport force that induces a small deviation in the direction of the radius
→ → →
CO. The differential equation of the relative motion of the material point, m a r = P + F C ,
is projected on the axes Ox and Oy and we find:
d 2x d 2y
m = F C , m = − mg (1)
dt2 dt2
The general solution of the second differential equation (1) is:
gt2
y (t) = H − (2)
2
The initial conditions t = 0 : x = 0, y = H were
considered. In addition,
dy
v r= v y= = − g t . In computing the magnitude of the Coriolis force it must be noted
dt

that the relative velocity v r of the point makes the angle 90 0 − λ with the axis of
rotation of the Earth, where λ is the latitude angle. In conclusion, FC = 2 m ω g t cos λ
and, from integration of the first equation (1), we find:
1
x ( t ) = ω g cos λ ⋅ t 3 (3)
3
Equations (2) and (3) are making up the set of parametric equations of the motion.
The motion is not rectilinear and the falling point shall have a deviation to the east. The
equation of the relative trajectory of the material point is obtained by eliminating the time
variable t between the two parametric equations. The semi-cubic parabola is found:
8ω2
x 2= cos 2 λ (H − y) 3 (4)
9 g
At the contact with the ground: y = 0. From (4) we find the deviation ε to the east
the point shall have with respect to the vertical line of the point of launch:
2 2H3
ε= ω cos λ (5)
3 g
Note that this expression for the deviation ε represents an approximate value, but
it has been validated by many experimental measurements.

13.5. Problems proposed

13.5.1. Conventional tests

TC 13.1) Two plane surfaces, between which a constant distance 2r is maintained, rotate
about a vertical axis at the constant angular velocity ω . Between the two surfaces moves

221
Theoretical mechanics – theory and applications


with friction a ball of mass m and radius r, launched at the initial instant at the velocity v 0
perpendicular to the axis of rotation (TC Figure 13.1). Determine the laws of motion
x=x(t), y=y(t), z=z(t) with respect to the chosen system of axes and the normal reaction
N=N(t).
z

y1
y
A M


 G
A V0
α
•M
B
x
• •

O1 x1

Figure TC 13.1 Figure TC 13.2.1

TC 13.2) The rod AB, inclined at an angle α with the horizontal direction, moves in the
horizontal direction in a rectilinear translation motion as per the law x 1 (t ) =
1
a t 2 . On the
2
rod, a material point M of mass m can slip without friction facing the environment
resistance proportional to the magnitude of the velocity R = m c v a , where v a is the
magnitude of the absolute velocity (Figure TC 13.2.1). At the initial instant the mobile M

is at the point A and has the velocity v 0 , horizontal (from A to B). Knowing the constants
α , a, m, c, v 0 , determine the law of the relative motion of the mobile with respect to the rod.

13.5.2. Multiple-choice tests

TG 13.1) The fundamental equation of the dynamics of the relative motion of the material
point is:
→ → → → → → → → → → →
a) m a a = F ; b) v a = v r + v t ; c) m a r = F + F t + F C ; d) 0 = F + F t .

TG 13.2) In the relative motion of a material point, the composition of the accelerations is
defined by the relation:
             
a) a a = a r + a t ; b) a a = a r + a t + a C ; c) a a = a C + a t ; d) a a = a r × a t × a C .

13.6. Indications and answers

TC 13.1) The relative motion of the ball M is a motion in the plane Oxz, characterized by
→ •→ • → → •• → •• →
the relative velocity v r = x i + y j and the relative acceleration a r = x i + y j . The
transport motion is a uniform circular motion at the angular velocity ω , on a circle of
radius equal to x (distance from point M to the radius Az). The differential equation of the
relative motion is:

222
Theoretical mechanics – theory and applications

→ → → → →
m a r = G+ N + F t + F C
→ → → → → → → → → • →
where G = − m g k , N = N j , F t = − m a t = m x ω 2
i şi F C = − m a C = −2 m ω x j .
Solving this equation in the initial conditions:
• •
t =0 : x= z =0, x=v0 , z =0
it is obtained:
gt2
sh (ω t ) , N = 2 m ω v 0 ch(ω t )
v0
x= y=0 , z= ,
ω 2
y
 
y1 j  N
A R

i M
  
Ft Vr , σ r

G
α x
B
O1
x1

Figure TC 13.2.2

TC 13.2) The Cartesian system Axy is considered as movable reference system (stiff-
connected to the rod AB) (Figure TC 13.2.2). The relative motion of the point M is
governed by the equation:
→ → → → → →
m a r = G+ N + R + F t + F C ,
→ → → → → → →
where G = m g sin α i − m g cos α j , N = N j , F C = 0 (because the transport motion is
the motion of a point of a body under translation),
→ → → →   •→ →
R = − m c v a = − m c  v r + v t  = − m c  a t cos α + x  i − m c a t sin α j .
   

The initial conditions of the motion are: t = 0 : x = 0 , x = v 0 . The motion of the
point M with respect to the rod is dictated by the law:
v 0 
x( t ) = 
g g
(
− 2 sin α  1 − e − c t + t sin α −
at2
)
cos α .
 c c  c 2

TG 13.1) Theoretical result. Correct answer: c).

TG 13.2) Theoretical result. Correct answer: b).

223
Theoretical mechanics – theory and applications

14. Mechanical moments of inertia


14.1. Definitions

The mechanical moments of inertia are scalar quantities that characterize the
spread of mass inside a material points system or a rigid solid.
By definition, the moments of inertia of a system of material points relative to a
plane, an axis or a point (pole) are the sum of the products of the masses of material points
system and squares distances from these points to the plane, axis or pole, respectively.
By definition, the centrifugal moments represent the sum of the products of masses
of the material points and the coordinates of these points in relation to two perpendicular
planes.

Let a system of n material points be related to a Cartesian reference system Oxyz,


located at points A i ( xi , y i , z i ), i = 1, n , and having the masses m i , i = 1, n (Figure T 14.1).
Moments of inertia are defined as follows:
a) Planar moments of inertia
n n n
J Oxy = ∑ m i z i2 , J O y z = ∑ m i x i2 , J O z z = ∑ m i y i2 (14.1)
i =1 i =1 i =1
b) Axial moments of inertia
( ) ( ) ( )
n n n
J x = ∑ m i y i2 + z 2
i , J y = ∑ m i z i2 + x i2 , J z = ∑ m i x i2 + y i2 (14.2)
i =1 i =1 i =1
c) Polar moments of inertia

∑ m i (x i2 + y i2 + z i2 )
n
J O= (14.3)
i =1
d) Centrifugal moments of inertia
n n n
J xy= ∑m i x i y i , J yz= ∑m i y i z i , J zx= ∑m i z i x i (14.4)
i =1 i =1 i =1

z z

Ai
zi • ( D)
z A(dm)

y y y
O yi O

xi
x
x
x

Figure T 14.1 Figure T 14.2

224
Theoretical mechanics – theory and applications

The planar, axial and polar moments of inertia are positive moments while the
centrifugal moments may be null, positive or negative. The unit for the moment of inertia
is kg ⋅ m 2 .
For a rigid solid (Figure T 14.2) the sums in the expressions (14.1 – 14.4) are
replaced by integrals. Thus we have:
e) Planar moments of inertia
J O x y = ∫D z 2 dm , J O y z = ∫D x 2 dm , J O z x = ∫D y 2 dm (14.5)
f) Axial moments of inertia
( ) ( )
J x = ∫D y 2 + z 2 dm , J y = ∫D z 2 + x 2 dm , J z = ∫D x 2 + y 2 dm ( ) (14.6)
g) Polar moments of inertia
(
J O = ∫D x 2 + y 2 + z 2 dm ) (14.7)
h) Centrifugal moments of inertia
J ∫
x y= D x y dm , J y z= D ∫ y z dm , J ∫
z x= D z x dm (14.8)

x, y, z represent the coordinates of an arbitrary infinitesimal element of mass dm of


the domain (D) occupied by the rigid.

14.2. Relationships between moments of inertia

Among the 10 moments of inertia defined above, there are the following relations
of dependence:
1
JO = J x+ J y+ J z
2
( )
(14.9)
1
( ) 1
( )
1
J O x y = J x + J y − J z , J O y z = J y + J z − J x , J O z x = J z + J x − J y (14.10)
2 2 2
( )
so that only six of them are independent ( J x , J y , J z , J x y , J y z , J z x ).

If the material point system is a plane system, located in the plane Oxy, then we
can note that:
J x z = J yz = 0 , J z = J x + J y (14.11)

so, merely the computation of the moments J x , J y and J x y is required.

14.3. Variation of the moments of inertia about parallel axes.


STEINER’s theorem

Let a system of n material points be A i , i = 1, n , of masses m i , i = 1, n , with the


center of mass at C. Let two parallel lines ( ∆ ) and ( ∆ 1 ) be so that C ∈ ∆ . Given the axial
moment of inertia J ∆ , let us compute, depending on it, the moment of inertia J ∆ 1 .

225
Theoretical mechanics – theory and applications

Therefor, we consider the following reference systems:


- the system Cxyz, for which ∆ ≡ C z . Related to this system, the point A i has the
coordinates x i , y i , z i , i = 1, n (Figure T 14.3);
- the system O 1 x 1 y 1 z1 , for which ∆ 1≡ O 1 z 1 and O 1 x 1 // C x, O 1 y1 // C y .
Related to this system the A i point has the coordinates x 1i , y 1i , z 1i . The obvious relations
take place:
x 1i = x i + x C , y 1i = y i + y C , z 1i = z i + z C (14.12)

(Δ )
z
z1 Ai z
zi y Ai
C (Δ )
xi di
yi
x
z1i
Bi
y1 O
O1 
uΔ y
x1i

x1 x
y1i

Figure T 14.3 Figure T 14.4

By definition,
( ) ( )
n n
J ∆ = J z = ∑ m i y i2 + z i2 , J ∆ 1 = J z 1 = ∑ m i y 21i + z 12i (14.13)
i =1 i =1

From (14.12) and (14.13) we obtain:

[
J ∆ 1 = ∑ m i ( x i + x C ) 2 + ( y i + yC ) 2 = ]
n

i =1

( ) ( )
n n n n
= x C2 + y C2 ⋅ ∑ m i + 2 x C ⋅ ∑ m i x i + 2 y C ⋅ ∑ m i y i + ∑ m i x i2 + y i2 (14.14)
i =1 i =1 i =1 i =1

n
Denoting the mass of the system by M = ∑ m i and the distance between axes
i =1

( ∆ ) and ( ∆ 1 ) by d = x C2 + y 2
C , noting that according to the theorem of the static

(as C (ξ )∈C z ,
n n
moments ∑ m i x i = M ⋅ ξ C = 0 , ∑m i y i = M ⋅η C= 0 C ,η C , ζ C from
i =1 i =1
(14.14) it is found that:
J∆1 = J∆ + M ⋅ d 2 (14.15)

Relation (14.15) represents the mathematical form of Steiner’s theorem. Its


statement is:

226
Theoretical mechanics – theory and applications

Steiner’s theorem: The moment of inertia about an axis ( ∆ 1 ) is equal to the sum
of the moment of inertia about an axis ( ∆ ), parallel to ( ∆ 1 ), containing the center of
mass, and the product of mass of the system and the square of the distance between the two
axes.

Customizing the relation (14.15) we find that:

J x1 = (
J x + M ⋅ y C2 + z C2 ) , J y1= (
J y + M ⋅ z C2 + x C2 ) , J z1= ( )
J z + M ⋅ x C2 + y C2 (14.16)

Using the relations of definition and (14.12) we can also find the connection
between the centrifugal moments computed relative to the two reference systems:

J x 1 y 1 = J x y + M x C y C , J y 1 z 1 = J y z + M y C z C , J z 1 x 1 = J z x + M z C xC (14.17)

14.4. Variation of the moments of inertia about concurrent axes

A system of n material points A i , i = 1, n is considered, of masses m i , i = 1, n , for


which the moment of inertia J x , J y , J z , J x y , J y z , J zx are known relative to the axes of
the Cartesian reference Oxyz.
The issue is to determine the moment of inertia related to an axis ( ∆ ) passing
through point O and having the direction given by the direction cosines cos α , cos β , cos γ
(Figure T 14.4).
The moment of inertia of the material point system about the axis ( ∆ ) is given by
the relation:
n n
J ∆ = ∑ m i d i2 = ∑ m i ⋅ A i B i2 (14.18)
i =1 i =1
d i = A i B i , being the distance from point A i to the line ( ∆ ) .
From ∆OA i B i we find that A i B i2 = OA i2 − OB i2 . But:
→  → → 2
OA i2 = r i2 = x i2 + y i2 + z i2 , OB i2 = OB i2 =  r i ⋅ u ∆  = ( x i cos α + y i cos β + z i cos γ ) 2
 
so that:

[( )( )
J ∆ = ∑ m i ⋅ x i2 + y i2 + z i2 cos 2 α + cos 2 β + cos 2 γ − ( x i cos α + y i cos β + z i cos γ ) 2 ]
since cos 2 α + cos 2 β + cos 2 γ = 1 . Upon convenient associations, from the preceding
relation it is obtained:
( ) ( ) ( )
n n n
J ∆ = cos 2 α ⋅ ∑ m i y i2 + z i2 + cos 2 β ⋅ ∑ m i z i2 + x i2 + cos 2 γ ⋅ ∑ m i x i2 + y i2 −
i =1 i =1 i =1
n n n
− 2 cos α cos β ⋅ ∑ m i x i y i − 2 cos β cos γ ⋅ ∑ m i y i z i − 2 cos γ cos α ⋅ ∑ m i z i xi ,
i =1 i =1 i =1

227
Theoretical mechanics – theory and applications

that is:
J∆ = J x cos 2 α + J y cos 2 β + J z cos 2 γ − 2 J x y cos α cos β −
− 2 J y z cos β cos γ − 2 J z x cos γ cos α (14.19)

Particular case: Should the points of the system and the axis ( ∆ ) belong to the plane Oxy,
π π
then, since z = 0, γ = ,α + β = , we find that:
2 2
J ∆ = J x cos 2 α + J y cos 2 β − 2 J x y sin α cos α (14.20)

14.5. Principal axes and moments of inertia

The moment of inertia J ∆ depends on the axis position related to the trihedral
Oxyz by means of the angles α , β , γ made by the axis ( ∆ ) with Ox, Oy and Oz,
respectively. Depending on the measure of these angles, J ∆ may have extreme values.
The axes ( ∆ ) passing through O, related to which the moments of inertia have extreme
values are called principal axes of inertia. The moments of inertia computed with respect
to these axes are called principal moments of inertia (we denote them by J 1 , J 2 , J 3 ).
These moments may be obtained by Lagrange multiplier method. Thus, let us
consider the function:
J ∆ ( cos α , cos β , cos γ ) = J x cos 2 α + J y cos 2 β + J z cos 2 γ − 2 J x y cos α cos β −
(
− 2 J y z cos β cos γ − 2 J z x cos γ cos α − λ cos 2 α + cos 2 β + cos 2 γ − 1 )
The extreme values are obtained as solutions of the system:

 ∂ J∆
 ∂ cos α = 2( J x − λ) cos α − 2 J x y cos β − 2 J zx cos γ = 0

 ∂ J∆

 ∂ cos β
( )
= −2 J x y cos α + 2 J y − λ cos β − 2 J y z cos γ = 0
(14.21)
 ∂J
= −2 J z x cos α − 2 J y z cos β + 2 ( J z − λ) cos γ = 0


 ∂ cos γ


The homogenous system (14.21), with the unknowns cos α , cos β , cos γ , has non
zero solutions only if its determinant is null:

J x − λ − J x y − J xz
− J yx J y − λ − J yz = 0 (14.22)
− J zx − J zy J z − λ

It is thus obtained a third degree equation with the unknown λ , which has three
real roots because the determinant is symmetrical with respect to the principal diagonal

228
Theoretical mechanics – theory and applications

(Kronecker’s theorem). The roots λ 1, λ 2 , λ 3 are even the extreme values of the
moment J ∆ , namely J 1, J 2, J 3 .
The principal axes of inertia ∆ i , i = 1,3 , have as direction parameters of their
direction the determinants:
J y − λi − J yz − J yz − Jyx − J yx J y − λi
, , .
− Jzy Jz − λi J z − λi − J zx − J zx − Jzy

Particular case: If the material point system is in the plane Oxy, then
since J z = J O = J x + J y , J x z = J y z = 0 , the equation (14. 22) has the solutions:

Jx + Jy  J x − J y 2
J 1, 2 = ±   + J 2x y , J 3 = J O (14.23)
2  2 

( )
and the direction parameters of the principal axes of inertia are:
- for ∆ 1: J y − J 1 ( J z − J 1 ) ; J x y ( J z − J 1 ) ; 0;
- (
for ∆ 2: J y − J 2 )( J z − J 2 ) ; J x y (J z − J 2 ) ; 0;
- ∆ 3≡ O z .

Note: The principal axes of inertia are forming a triorthogonal trihedral and the
centrifugal moments with respect to the axes of this trihedral are null.
If the center of mass of the system coincides with the origin of the
trihedral ( O ≡ C ) , then the moments of inertia corresponding to the axes passing through C
are called central moments of inertia. The moments about the principal axes of inertia with
respect to the center of mass are called central principal moments of inertia.

14.6. Solved problems

R 14.1) Let us consider the homogenous plate in Figure R 14.1.1 for which the mass M
and the distance a are known. Requested:
a) The axial moments of inertia J x , J y and the centrifugal moment J x y ;
b) The principal moments and axes of inertia
y
y

2a
6a b O
2a x

O 2a x
a

Figure R 14.1.1 Figure R 14.1.2

229
Theoretical mechanics – theory and applications

Solution: Assume as known the moments of inertia for a homogenous rectangular


flat plate of mass m and sides a and b (Figure R 14.1.3):

J O= J z=
(
m a 2+ b 2
, J x=
)mb 2
, J y=
ma2
, J xy = J yz = J zx = 0
12 12 12
The plate area is A = A 1+ A 2 = 6a ⋅ 2a + 2a ⋅ 2a = 16 a 2 and the
M M
density ρ = = . The masses of the two rectangular plates the plate has been
A 16 a 2
3M M
divided into are m 1= ρ A 1= and m 2 = ρ A 2 = .
4 4
a) J x = J 1 x + J 2 x , J y = J 1 y + J 2 y , J x y = J 1 x y + J 2 x y (1)
where:
m 1( 6 a ) 2
J 1x = J 1 x 1 C + m 1⋅ ( 3a) 2 = + m 1( 3a) 2 = 9 M a 2
12
m 2 ( 2a ) 2 19
J 2 x = J 2 x 2 C + m 2 ⋅ ( 5a ) 2 = + m 2 ( 5a ) 2 = M a 2
12 3
(
m 1 2a ) 2
J 1 y = J 1 y 1 C + m 1⋅ ( a) 2 = + m 1( a ) 2 = M a 2 (2)
12
m 2 ( 2a ) 2 7
J 2 y = J 2 y 2 C + m 2 ⋅ ( 3a) = 2
+ m 2 ( 3a) 2 = M a 2
12 3
9
J 1 x y = J 1 x 1 C y 1 C + m 1⋅ a ⋅ 3a = 0 + m 1⋅ a ⋅ 3a = M a 2
4
15
J 2 x y = J 2 x 2 C y 2 C + m 2 ⋅ 3a ⋅ 5a = 0 + m 2 ⋅ 3a ⋅ 5a = M a2.
4

In finding of the relations (2) Steiner’s formulas have been used. From (1) and (2)
we find that:
y y y1
y 2C

2 2a
2a
y1C C 2 (3a ,3a ) x 2C

2a 6a
1 2a
6a
C1 (a ,3a ) x1C

α2
x
O α1
⋅ 2a
O 2a x z
x1

Figure R 14.1.3 Figure R 14.1.4

230
Theoretical mechanics – theory and applications

46 10
Jx = M a2 , J y= M a2 , Jxy = 6 M a2 (3)
3 3

b) In the particular case of a plane material system (z = 0) the principal moments of inertia
are given by the relations:
J x+ J Jx − Jy 2
±   + J
y
J 1, 2 = 2
xy , J 3= J z = J x + J y = J1 + J 2 (4)
2  2 
The values obtained are:
J 1= 17,91 M a 2 , J 2 = 0,75 M a 2 , J 3 = 18,56 M a 2 (5)
The principal axes of inertia (Figure R 14.1.4) are forming with the axis Ox the
angles α 1, α 2 and α 3 given by the relations:
J x− J 1 Jx − J 2
tg α 1= = −0,614 , tg α 2 = = 2,414
Jxy Jxy
π
α 3= (the third principal axis of inertia coincides with Oz).
2

R 14.2) For the system of bodies in Figure R 14.2.1, formed by two homogenous discs of
radius R and mass M and four homogenous bars of length l and mass m, determine the
axial moments of inertia J x , J y , J z and the polar moment of inertia J O .

R y
y

l /2
l /2 l /2 M, R
l x
O O
O y x
O m, l ⋅
⋅ z
z
x

Figure R 14.2.1 Figure R 14.2.2

Solution: Let us assume as known the moments of inertia for a homogenous bar of
mass m and length l, and for a homogenous disc of mass M and radius R, respectively
(Figure R 14.2.2):
ml 2 ml 2
JO = ,Jx = 0,Jy = Jz = , J xy = J yz = J zx = 0 ,
12 12
M R2 MR2
JO = J z= ,Jx = Jy = , J xy = J yz = J zx = 0 .
2 4

231
Theoretical mechanics – theory and applications

6
For the body in Figure R 14.2.1 we have that J x = ∑ J i x , where:
i =1

ml 2
ml 2
MR 2  l 2
J 1x = J 2 x = , J 3x = J 4 x = + mR 2 , J 5 x = J 6 x = + M  .
12 12 4 2
It results:
M  M m
J x =  + 2m  R 2 +  +  l 2 (1)
 2   2 3
Owing to the symmetry we may write that:
J y= J x (2)
6
In the same time, J z = ∑ J i z , where:
i =1

M R2
J 1z = J 2 z = J 3 z = J 4 z = 0 + m R 2 = m R 2 , J 5 z = J 6 z = .
2
We obtain hence that:
J z = (M + 4 m) R 2 (3)
Eventually, the polar moment of inertia has the value:
J x+ J y+ J z M m
J O= = (M + 4 m) R 2 +  +  l 2
(4)
2  2 3

R 14.3) Given a homogenous cone shaped plate of mass M, radius of the base R and height
h (Figure R 14.3.1). Requested:
a) The moments of inertia about the axes of the coordinate system with the origin
at the vertex of the cone and having the cone axis as axis Oz;
b) Write the matrix of the moments of inertia;
c) The moments of inertia about the axes of the coordinate system with the origin
at the center C of the masses, and having the cone axis as Cz axis. Write the
matrix of the moments of inertia with respect to this system.

z, z c z

R A

C B

yc
A′ r
xc h
B′
z l

O
O y y
x θ
x

Figure R 14.3.1 Figure R 14.3.2

232
Theoretical mechanics – theory and applications

Solution: a) From symmetry reasons we have J x = J y , so that only the moments


of inertia J x and J z shall be determined, given by the relations:
( )
J x = ρ ∫D y 2 + z 2 dA , J z = ρ ∫D x 2 + y 2 dA ( ) (1)
M M
where ρ = = represents the superficial density.
A π RG
For the computation of the two integrals the conical surface is divided into
infinitesimal elements (Figure R 14.3.2). The position of such an element is given by the
coordinates:
x = r cos θ , y = r sin θ , z = z (2)
and the element of area by dA = r dl dθ . Since ∆OA' B' ≈ ∆OAB we may write
R h G
that = = , thus:
r z l
R G G RG
r = z , l = z , dl = dz , dA = 2 z dz dθ (3)
h h h h
Substituting these relations in the formulas (1) it is obtained:
ρ RG R2 
Jx = ∫∫  z 2
sin 2
θ + z 2
 z dz dθ =
h2 D h2
 
ρR G  R 2 h 2π
R2 h 2π h 2π 
=  ∫ z 3 dz ⋅ ∫ dθ − 2 ∫
z 3 dz ⋅ ∫ cos 2θ dθ + ∫ z 3 dz ⋅ ∫ dθ  =
h2 2 h
2
0 0 2h 0 0 0 0 
M 2
4
=
R + 2h 2 . ( ) (4)

ρR G3
ρ R 3G h 3 M R2
J z= ∫∫D z dz dθ = 4 ∫ z dz ⋅ ∫ dθ = 2 .
3
h4 h 0 0
b) From symmetry reasons we may state that:
J x y = J y z = J zx = 0 (5)
so that the matrix of the moments of inertia is:

(
 M R 2 + 2h 2 )0 0


 4 
[J] =

0
M R + 2h 2
2
( ) 0 

(6)
4
 
 M R2 
2 
0 0

c) Steiner’s formula is used and it is obtained that:

JxC = J x − M ⋅ OC = 2
(
M R 2+ 2 h 2
−M⋅
4 h2 M
=
)  2
 R 2 + 2 h  (7)
4 9 4  9 
 2 2
J y C = J y − M ⋅ OC 2 =
M  R 2 + 2 h , J = J − M ⋅ 0 2 = M R .
4  9  zC z
2

233
Theoretical mechanics – theory and applications

From symmetry reasons:


J x C yC= JyC zC = J zC xC = 0 (8)

The matrix of the moments of inertia remains diagonal:



(
 M R 2 + 2h 2 / 9
0
)
0 

 4 

[J C ] =  0
M R 2 + 2h 2 / 9
0 
 ( ) (9)
4
 
 M R2 
2 
0 0

14.7. Problems proposed

14.7.1. Conventional tests

TC 14.1) Let us consider the material points system in Figure TC 14.1, placed at points of
the surface of the rectangular parallelepiped [ABCDA’B’C’D’] of sides AB = AD = 2a
and AA ' = 4a . Knowing the mass m, determine:

a) The axial J x , J y , J z and centrifugal J x y, J yz , J z z moments of inertia;


b) The polar moment of inertia JO and the planar moments of
inertia J O x y , J O y z , J O z x .
z
r + dr r
z

D′ m
C′ •
3

B′
A′ 4m H
2
1 C
2m D
4

O 2m y
O
y
A
B R
x x
(Δ )
Figure TC 14.1 Figure TC 14.2

TC 14.2) Determine the moment of inertia for a homogenous right circular cylinder of
mass M, radius of the base R and height H, about its axis of symmetry (Figure TC 14.2).

TC 14.3) Compute the moments of inertia of a homogenous sphere of radius R and mass M
about the planes of symmetry and the point O (Figure TC 14.3).

234
Theoretical mechanics – theory and applications

x
y

y l
x x + dx
R x dx
O z x

A B

y
z

Figure TC 14.3 Figure TG 14.1

14.7.2. Multiple-choice tests

TG 14.1) Determine the moment of inertia of a straight bar AB, of length L, of which
density varies linearly from the value ρ 1 (at the end A) to the value ρ 2 (at the end B),
about the center C of the bar.
(3 ρ 2 + ρ 1 ) L 3 ( ρ 2+ 3 ρ 1 ) L 3 (3 ρ 2 − ρ 1 ) L 3
a) J C = ; b) J C = ; c) J C = ;
12 6 24
( ρ 2− 3 ρ 1 ) L 3
d) J C =
12

TG 14.2) What is the value of the polar moment of inertia J O of a homogenous flat plate
of circular sector shape of radius R, angle at center 2 α and superficial density ρ .
ρ R 4α ρ R 4α π ρ R 4α
a) J O= ; b) J O = ; c) J O = 2 ρ R 2 α ; d) J O = .
6 2 12

14.8. Indications and answers

TC 14.1) a) Formulas (14.2) and (14.4) from section 14.1 are applicable. The computation
is made with the help of the table below showing the masses and the coordinates of the
four material points.

Item no. mi xi yi zi
1 2m 0 -a 2a
2 4m a a 4a
3 m -a 0 4a
4 2m 0 a 0

The following values are obtained:


J x = 13 m a 2 , J y = 93 m a 2 , J z = 96 m a 2

x y= 4 m a2 y z = 12 m a z x = 12 m
2
J , J , J a2.

235
Theoretical mechanics – theory and applications

b) For obtaining the polar moment of inertia J O and the planar moments of
inertia J O x y , J O y z , J O z x , formulas (14.1) and (14.3) are used. It results:
J O = 101 m a 2
, J Ox y = 5 m a 2 , J O y z = 88 m a 2 , J O zx= 8 m a 2 .

TC 14.2) The cylinder is divided into infinitesimal elements of the type in Figure TC 14.2
and it is obtained:
R
π H ρ R 4 M R2
J ∆ = ∫ r 2 dm = ∫ r 2 ρ ⋅ 2 π r l dr = = ,
0 2 2

as M = ρ V = ρ ⋅ π R 2 H .
TC 14.3) A division of the sphere into infinitesimal elements is considered as in Figure TC
14.3. Owing to symmetry we may write that J O x y = J O y z = J O z x . Yet:

∫D x ∫D x ∫
2 2 M
JO yz = dm = ρ dV = x 2 dV , and
2V D
2
dV = π y 2dx , y 2 = R 2 − x 2, V = πR 3 ,
3
1
so that J O y z = M R 2 . The polar moment of inertia of the sphere equals to:
5
( ) 3
J O = ∫D x 2 + y 2 + z 2 dm = J O x y + J O y z + J O z x = M R 2 .
5
ρ 2− ρ 1
TG 14.1) The law of density variation in the bar is ρ ( x) = ⋅ x + ρ 1 . Since on the
L
bar y = z = 0, from (14.7) it follows:
L
 ρ 2− ρ 1  (3 ρ 2 + ρ 1 ) L 3
J A = ∫AB x 2 dm = ∫AB x 2 ρ dx = ∫  ⋅ x + ρ 1  ⋅ x 2 dx =
0 L  12
From Steiner’s formula we obtain:
(3 ρ 2 − ρ 1 ) L 3
J C = J A − M ⋅ AC 2 = ,
24
L L ( ρ 1+ ρ 2 ) L
since AC = and M = ∫AB dm = ∫ ρ ( x) dx = . Correct answer: c).
2 0 2

TG 14.2) Since the plate is in the plane Oxy (z = 0), from (14.7) we find that:
( )
J O = ∫D x 2 + y 2 dm = ρ ⋅ ∫D x 2 + y 2 dA . ( )
The polar coordinate system ( r , θ ) may be used, for which:
x = r cos θ , y = r sin θ , dA = r dr dθ
It is obtained:
α
R
ρ R 4α
J O = ρ ⋅ ∫ r 3 dr ⋅ ∫ d θ = .
0 −α 2
Correct answer: b).

236
Theoretical mechanics – theory and applications

15. General theorems in the dynamics of the systems of material


points and of rigid solids
15.1. Linear momentum theorem

15.1.1. Linear momentum of a system of material points and of a rigid solid


The linear momentum of a material point of mass m, of velocity v , is a vector

collinear and of the same sense with the velocity v , defined by the equation:
→ def →
H = mv (15.1)

For a system of material points of masses m i , i = 1, n , and velocities v i , i = 1, n , the
linear momentum of the system is equal to the vector sum of the moments of the material
points the system is made up from:
→ def n →
H = ∑m i vi (15.2)
i =1
The linear momentum of a rigid solid is defined by the equation:
→ def →
H = ∫D v dm (15.3)
where the integral is considered on the entire domain (D) occupied by the rigid.

Notes: i) The SI unit for the linear momentum is kg ⋅ m ⋅ s −1 .


ii) Taking into account the equation of definition of the position of the center of
n →


∑m i r i
i =1
mass r C = n
, and the definition (15.2), we may write:
∑m i
i =1

→ n d ri d n →  d  →  →
H = ∑ m i⋅ = ∑ m i r i  = M r C  = M v C .
i =1 dt d t  i =1  d t 
Expression
→ →
H =M vC (15.4)
n →
where M = ∑ m i is the mass and v C the velocity of the center of mass, shows that the
i =1
linear momentum of a system of material points may be considered as linear momentum of
the center of mass wherein the whole mass of the system is supposedly concentrated.

237
Theoretical mechanics – theory and applications

iii) The demonstrations in this chapter are made for the case of the system of material
points but they remain valid for the case of the rigid solid as well.

15.1.2. Linear momentum theorem. Theorem of the motion of the center of mass

Linear momentum theorem (statement): The derivative with respect to time of


the linear momentum of a system of material points or of a rigid equals to the sum of the
external forces acting on the system or the rigid:

→ →
H = ∑ F ext (15.5)

Proof: Let us consider a system of n material points of masses m i , i = 1, n , and



accelerations a i , i = 1, n , found in interaction (Figure T 15.1). The point A i is acted upon
by two categories of forces:

- External forces, replaced by their resultant F i ;

- Internal forces, F i j , i, j = 1, n, j ≠ i , namely the forces with which the other
points of the system act upon the point A i . According to the principle of action and
reaction the internal forces are in pair and directly opposite, that is:
→ → → → → → → →
Fi j+ F ji =0 , r i× Fi j+ r j× F ji =0 (15.6)
Singling out the system points and writing for each the fundamental equation of
the dynamics, the following equations are obtained:
→ → → → → →
m i a i = F i + F i 1 + F i 2 + ... + F i j + ... + F i n , i = 1, n (15.7)
Summating member by member the n equations (15.7) we find that:
n → n → n n →
∑m i a i = ∑ F i + ∑ ∑ F i j (15.8)
i =1 i =1 i =1 j =1
j ≠i
Based on the first equation (15.6), the final sum in (15.8) is null. In addition:
→ → •
n → dv d n n → dH →
∑ m i a i = ∑ m i d t i = dt ∑m i v i = d t = H (15.9)
i =1 i =1 i =1
thus:

→ n → →
H = ∑ F i = ∑ F ext (15.10)
i =1

In terms of projections on the axes of the Cartesian reference Oxyz, the equation
(15.10) is written:
• n • n • n
H x = ∑ Fi x , H y = ∑ Fi y , H z = ∑ Fi z (15.11)
i =1 i =1 i =1

238
Theoretical mechanics – theory and applications

Theorem of the motion of the center of mass (statement): The center of mass of
a system of material points or of a rigid moves similarly to a point wherein the whole mass
of the system (rigid) is concentrated and whereupon all the external forces are acting:
→ →
M a C = ∑ F ext (15.12)
→ →
Proof: Differentiating the relation H = M v C with respect to time t we find
• •
→ → → → →
that H = M v C = M a C . From (15.10) we find then that M a C = ∑ F ext .

Note: The theorem of linear momentum and the theorem of motion of the center of mass
are not independent theorems. The theorem of motion of the center of mass represents
another form of presentation of the theorem of linear momentum.

15.1.3. Theorem of linear momentum conservation

Theorem of linear momentum conservation (statement): If during the motion


the system of material points (the rigid) is isolated or if the sum of the external forces is
null, then the linear momentum of the system (of the rigid) is conserved.

→ → → → → →
Proof: Since ∑ F ext = 0 , from (15.5) we find that H = 0 , i.e. H = M v C =
constant.

Note: In many practical cases the resultant of the external forces only has null the
component about a single axis, which will lead to the conservation of the linear momentum
about that axis only. If such an axis is Ox, then:

∑ Fx, ext = 0 ⇒ H x = M ⋅ vC x = constant (15.14)

zc
Ai   Aj
Fij Fji Ai
 z
Fi 
  r′i
Fj ri
  C
r1i r1 j
xc yc

rc
O1 O1 y
x

Figure T 15.1 Figure T 15.2

239
Theoretical mechanics – theory and applications

15.2. Theorem of the angular momentum

15.2.1. The angular momentum of a system of material points and of a rigid solid.
Koenig’s Theorem for the angular momentum

The angular momentum of a material point of mass m, moving at the velocity v ,
→ →
computed about a fixed point O, is the moment of the momentum vector H = m v of the
point computed about the point O:
→ def → → → →
KO = r× H = r×m v (15.15)

The angular momentum of a system of material points of masses m i , i = 1, n , and



velocities v i , i = 1, n , about a fixed point O, is obtained with the relation:
→ def n → n → →
KO = ∑ K Oi = ∑ r i × m i vi (15.16)
i =1 i =1
The angular momentum of a rigid solid is defined by the relation:
→ def → →
KO = ∫D r × v dm (15.17)
where the integral extends over the entire domain (D) occupied by the rigid.

Koenig’s theorem for the angular momentum (statement): The angular


momentum of a system of material points or a rigid about a fixed point O 1 is equal to the
sum of the angular momentum about O 1 of the center of mass wherein it is assumed the
entire mass of the system (rigid) is concentrated and the angular momentum of the system
(rigid) in its relative motion about the center of mass:
→ → → →
K O1 = r C × M v C + K C (15.18)
Proof: Let us consider the system of material points A i , i = 1, n , of
masses m i , i = 1, n . It is related to the Cartesian system O 1 xyz and to a Cartesian system
C xC y C z C with the origin at center of mass and the axes parallel to those of the system
O 1 xyz (Figure T 15.2). Between the position vectors of the point A i with respect to the
two references there is the relation:
→ → →
r i = r C + r i' (15.19)
By differentiation with respect to time, from (15.19) it is obtained:
→ → →
v i = v C + v i' (15.20)
From (15.16), (15.19) and (15.20) we find that:
→ n → →  → →  n → → n → →
K O 1 = ∑  r C + r i' × m i  v C + v i'  = ∑ r C × m i v C + ∑ r C × mi v i' +
i =1     i =1 i =1
n → → n → →
+ ∑ r i' × m i v C + ∑ r i' × mi v '
i
i =1 i =1

240
Theoretical mechanics – theory and applications

But:
n → → →  n → → →
∑ r C × m i v C = r C × ∑ m i  v C = r C × M vC
i =1  i =1 
n → → → d  n →  →
∑ r C × m i v i' = r C × d t  ∑ m i r i'  = 0 (15.21)
i =1  i =1 
→ → n →  → → n → → →
∑ r i' × m i v C =  ∑ m i r i'  × v C = 0 , ∑ r i × m i v i' = K C
 i =1  i =1
n → → →
The fact that ∑m i r i' = S C = M ⋅ ξ C is the polar static moment about the
i =1
→ →
center of mass was taken into account. It is null as ξ C = 0 (the position vector of the
point C with respect to the reference C xC y C z C ).
From (15.20) and (15.21) we find now (15.18).

15.2.2. Expressions of the angular momentum in different particular motions


of the rigid

15.2.2.1. Motion of translation

In the motion of translation the velocity of all points of the rigid are, at an instant,
equal to each other and equal to the velocity of the center of mass. Since there is no
relative motion about the center of mass, we have:
→ → → → →
KC = 0 , K O1 = r C × M v C (15.22)

The angular momentum of the rigid about the point O 1 , in the translation motion,
is obtained as if the entire mass of the rigid would be concentrated in the center of mass
and would move at the velocity thereof.

15.2.2.2. Motion of rotation

In the motion of rotation the velocity of a point A i is obtained by the


→ → →
relation v i = ω × r i . We shall take the general case wherein the axis of rotation has some
direction related to a movable trihedral stiff-connected to the rigid. The angular velocity
→ →
vector ω and the position vector r i are given by:
→ → → → → → → →
ω = ω x i + ω y j + ωz k , r i = xi i + yi j+ zi k
We have in order:

241
Theoretical mechanics – theory and applications

→ → →
i j k
→ → →
v i = ω× r i = ω x ωy ( ) →
ω z = ω y z i − ω z y i i + (ω z x i − ω x z i ) j + ω x y i − ω y x i k

( ) →

xi yi zi
→ → →
i j k
→ n → → n
K O1 = ∑ r i × m i v i = ∑ mi xi m i yi mi zi
i =1 i =1
ω y zi − ω z y i ω z x i− ω x z i ω x y i− ω y x i

Projection K x of the angular momentum K O 1 on O 1 x shall be:
mi yi mizi
[ ( ) ]
n
Kx = ∑ = ∑ mi y i2 + z i2 ω x − mi x i y i ω y − m i x i z i ω z =
i =1 ω z x i − ω x z i ω x y i− ωy x i
= J x ω x− J x y ω y− J xz ω z .
Proceeding similarly in the directions O 1 y and O 1 z we obtain the following
projections of the angular momentum in the rotation motion:
K x= J x ω x− J x y ω y− J xz ω z , K y= −J yx ω x+ J y ω y− J yz ω z
K z= −J zx ω x− J z y ω y+ J z ω z (15.23)

Particular cases:
1) If axis O 1 z coincides with the axis of rotation, then ω x = ω y = 0, ω z = ω . The
projections of the angular momentum become:
K x= −J xz ω , K y= −J yz ω , K z = J z ω (15.24)
2) If axis O 1 z coincides with the axis of rotation and the rigid is a body of revolution
( )
J x z = J y z = 0 , then the angular momentum has the direction of the axis of rotation:
→ → → →
K O1 = K z k = Jz ω k = Jz ω (15.25)

15.2.2.3. Motion of the rigid with fixed point

In the motion of the rigid with fixed point (spherical motion) the velocity of an
arbitrary point A i has the same form as the one in the motion of rotation, so that the
expressions (15.23) of the projections of the angular momentum remain valid. But the

vector ω does no longer keep the fixed line of action.
If the axes of the movable reference Oxyz are chosen to coincide with the principal
axes of inertia related to the fixed point O 1 , then:
J x y = J y z = J zx = 0 , J x = J 1 , J y = J 2 , J z = J 3
and
→ → → →
K O = J 1ω x i + J 2 ω y j + J3 ω z k (15.26)

242
Theoretical mechanics – theory and applications

J1 , J 2 and J 3 are the principal moments of inertia about O1 .

15.2.2.4. Helical motion and plane-parallel motion

As in this particular motions the distribution of velocities is obtained by vector


summation of the of two distributions of velocities, one corresponding to a translational
motion and the other to a motion of rotation, the angular momentum related to the origin of
the fixed reference is obtained suchlike :
→ → →
K O 1 = K O 1 , translatie + K O 1 ,rotatie (15.27)

15.2.3. Theorem of the angular momentum in the motion of a system of


material points or a rigid related to a fixed reference

The theorem of the angular momentum (statement): The derivative with


respect to time of the angular momentum of a system of material points (rigid) computed
about a fixed point O1 is equal to the sum of the moments of the external forces acting
upon the system (rigid), moments computed about the same fixed point:

→ →
K O 1 = ∑ M O 1 ,ext (15.28)
Proof: Let us consider the system of material points, studied in paragraph 15.1.2.

Multiply vectorially at the left the relations (15.7) by r 1i , i = 1, n , and summate the
relations thus obtained:
n → → n → → n n → →
∑ r 1i × m i a i = ∑ r 1i × F i + ∑ ∑ r 1i × F i j (15.29)
i =1 i =1 i =1 j =1
j ≡1
Based on the second relation (15.6) the double sum in (15.29) is null. The first sum
in the right member is even the resultant moment of the external forces computed about O1 .
In addition,
→  → 
n → → n → n  → →  d r 1i →
∑ r 1i × m i a i = ∑ r 1i × m i dt = ∑ d t  r 1i × m i v i  − d t × m i v i  =
 
d vi d
i =1 i =1 i =1
 

d n → →  n → → d K O1
=  ∑ r 1i × m i v i  − ∑ v i × m i v i = .
d t  i =1  i =1 dt
→ → 
the second sum being null  v i // m i v i  .
 
In terms of projections on the axes of the fixed Cartesian reference O 1 x y z the
relation (15.28) is written:
• • •
K x = ∑ M O 1 x, ext , K y = ∑ M O 1 y , ext , K z = ∑ M O 1 z , ext (15.30)

243
Theoretical mechanics – theory and applications

15.2.4. Theorem of the angular momentum in the motion of a system of


material points or a rigid around the center of mass

Statement: The derivative with respect to time of the angular momentum of a


system of material points (rigid), corresponding to its motion around the center of mass is
equal to the resultant moment about the center of mass of the external forces acting on the
system:

→ →
K C = ∑ M C ,ext (15.31)

Proof: The relations (15.18), (15.19) and (15.28) are used.


→ → d → → →  n → →  →
K O 1 = ∑ M O 1, ext ⇒  r C × M v C + K C  = ∑  r C + r i'  × F i ⇒
dt  i =1  
→ → •
drC → → d vC → → n → n → →
⇐ × M v C + r C× M + K C = r C × ∑ F i + ∑ r i' × F i ⇒
dt dt i =1 i =1
 

=0
• •
→ → → → → → → →
⇒ r C × M a C + K C = r C × M a C + ∑ M C , ext ⇒ K C = ∑ M C , ext .

→ n →
The fact was used that M a C = ∑ F i (the theorem of the motion of the center of mass).
i =1

15.2.5. Theorem of angular momentum conservation

The theorem of angular momentum conservation (statement): If during the


motion, the system of material points (rigid) is isolated or if the resultant moment of the
external forces about O 1 is null, then the angular momentum of the system (rigid) about
the point O 1 is conserving.

→ → → →
Proof: As ∑ M O 1, ext = 0 , from (15.28) it is obtained that K O 1 = 0 , i.e.

K O 1 = constant.

Note: If only one of the projections of the resulting moment of the external forces is null,
then the angular momentum is only conserving about that axis. Thus:

M O 1 x ,ext = 0 ⇒ K O 1x = constant.

244
Theoretical mechanics – theory and applications

15.3. Work-Kinetic energy theorem

15.3.1. Kinetic energy of a system of material points and of a rigid solid.


Koenig’s theorem for kinetic energy


By definition, the kinetic energy of material point of mass m and velocity v is
given by the equation:
def 1
E = mv2 (14.32)
2

For a system of material points of masses m i , i = 1, n , and velocities v i , i = 1, n , the
kinetic energy is defined by the equation:
def n 1
E = ∑ m i v i2 (14.33)
i =1 2

The kinetic energy of a solid rigid is defined by the equation:

def 1 2
E = ∫D 2
v dm (14.34)

where the integral is considered over the entire domain occupied thereby.

Koenig’s theorem for the kinetic energy (statement): The kinetic energy of a
system of material points (rigid) in motion is equal to the sum of the kinetic energy of the
mass center wherein the whole mass of the system (rigid) is presumably concentrated and
the kinetic energy in the relative motion of the system (rigid) around the center of mass.
1
E= M v C2 + E ' (14.35)
2

Proof: Let us consider the system of material points studied in the paragraph
15.2.1. (See also Figure T 15.1). Using the definition (15.33) and equation (15.20) we find
that:

n 1 n 1 → n 1 → → 2
E=∑ m i v i2 = ∑ m i v i2 = ∑ m i  v C + v i'  =
i =1 2 i =1 2 i =1 2  
n 1 → → → →  1 n → n → n 1
= ∑ m i  v C2 + 2 v C v i' + v i  = v ∑m i + v C ∑m i v +∑
'2 2 ' '2
mi vi .
i =1 2   2 C
i =1 i =1
i
i =1 2
n n → → n 1
Noting that ∑ m i = M (mass of system), ∑m i v '
i = 0 and that ∑ 2 m i v i' 2 = E ' (the
i =1 i =1 i =1
kinetic energy in the relative motion of the system around the center of mass), the equation
(15.35) is obtained.

245
Theoretical mechanics – theory and applications

15.3.2. Elementary work and total work of a force



By definition, the elementary work performed by the force F is equal to:
 ∧ 
def → → → → →→
d L = F ⋅ d r = F ⋅ d r ⋅ cos F , d r  (15.36)
 
 

where d r is the elementary displacement of the point of application of the force.
→ →
Using the analytical expression of the vectors F and d r , depending on their
projections on the axes of a Cartesian reference Oxyz, we may write that:
d L = F x ⋅ dx + F y ⋅ dy + Fz ⋅ dz (15.37)

The total work corresponding to a variable force F and a total displacement
between two positions, A and B, respectively is defined by the relation:
def → →
L AB = ∫AB F ⋅ d r = ∫AB Fx ⋅ dx + F y ⋅ dy + Fz ⋅ dz (15.38)

The work corresponding to a couple of forces of moment M and to an angular
displacement Θ ∈ [Θ 1,Θ 2 ] is defined by the relation:
 ∧ 
def Θ → → Θ2 → →  → →
L AB = ∫Θ1
2
M ⋅ d Θ = ∫Θ
1
M ⋅ d Θ ⋅ cos M , d Θ 
 
(15.39)
 

15.3.3. Forms of the kinetic energy in various particular motions of the rigid

15.3.3.1. Translational motion

Since all points have the same velocity at an instant, it follows that:
n 1 n 1 1 n 1
E = ∑ m i v i2 = ∑ m i v C2 = v C2 ⋅ ∑ m i = M v C2 (15.40)
i =1 2 i =1 2 2 i =1 2
n →
where M = ∑ m i is the mass of the rigid and v C the velocity of center of mass.
i =1

15.3.3.2. Rotation motion


In a rotation motion all the points move at the same angular velocity ω and the
magnitude of velocity of the point A i is v i = l i ω , where l i is the distance from the point
to the axis of rotation. From (15. 33) we find that:
n 1 n 1 1 n 1
E = ∑ m i v i2 = ∑ m i l i2 ω 2 = ω 2 ∑ m i l i2 = J ∆ ω 2 (15.41)
i =1 2 i =1 2 2 i =1 2

246
Theoretical mechanics – theory and applications

where J ∆ is the moment of inertia about the axis of rotation ∆ .

15.3.3.3. Fixed point rigid motion

Components of the velocities of the point A i ( xi , y i , z i ) are:


vi x = z i ω y − y i ω z , vi y = x i ω z − z i ω x , vi z = y i ω x − x i ω y .
Noting that v i2 = v i2x + v i2y + v i2z , from (15.33) it follows that:

[( ) + (x ω (
− z iω x )2 + y iω x − x iω y ) ]=
n 1 1 n
E=∑ m i v i2 = ∑ m i z i ω y − y i ω z 2
i z
2
i =1 2 2 i =1

∑ m i (y i2 + z i2 ) + 2 ω 2y ∑ m i (z i2 + x i2 ) + 2 ω 2z ∑ m i (x i2 + y i2 ) −
1 n 1 n 1 n
= ω 2
x
2 i =1 i =1 i =1
n n n
− ω y ω z ∑ m i y i z i − ω z ω x ∑ m i z i xi − ω x ω y ∑ m i x i y i .
i =1 i =1 i =1
Thus:
1 1 1
E = J x ω 2x + J y ω 2y + J ω 2 − J y z ω y ω z − J z x ω z ω x − J x y ω x ω y (15.42)
2 2 2 z z

inertia (J )
If the axes of the movable reference Oxyz (O fixed point) are principal axes of
x y = J y z = J z x = 0 , J x = J 1, J y = J 2 , J z = J 3 , then:
1 1 1
E= J 1 ω 2x + J 2 ω 2y + J 3 ω 2z (15.43)
2 2 2

15.3.3.4. Helical motion

Components of the velocity vector of the point A i ( xi , y i , z i ) on the axes of the


movable reference Oxyz are:
v i x = −ω y i , v i y = ω x i , vi z = vO
So that:
[ ( ) ]
n 1 n 1
E=∑ m i v i2 = ∑ m i ω 2 x i2 + y i2 + v O
2
=
i =1 2 i =1 2
1 2n
( )
1 n 1 1
ω ∑ m i x i2 + y i2 + v O2 ∑ m i = J ∆ ω 2 + M v O2
= (15.44)
2 i =1 2 i =1 2 2
where J ∆ is the moment of inertia of the rigid about the axis of the helical motion and M is
the rigid mass.
15.3.3.5. Plane – parallel motion

The Cartesian components of the velocity vector of the point A i ( xi , y i , z i ) on the


axes of a movable trihedral, stiff-connected to the rigid, and having the plane Oxy parallel
to the fixed plane whereto the motion is related and the origin O even at the center of
masses C, are:

247
Theoretical mechanics – theory and applications

v i x = vO x − ω y i , v i y = vO y + ω x i , v iz= 0
So that:
 ( ) + (v ) 2
n 1 n 1
E=∑ m i v i2 = ∑ m i  v C x − ω y i
2
C y+ ω xi  =
i =1 2 i =1 2

( )∑ m ( )
1 2 n 1 2n
= v +v2 + ω ∑ m i x i2 + y i2 −
2 Cx Cy i =1
i
2 i =1
n 1 n 1
− ω vC x ∑ m i y i − ω vC y ∑ m i xi ⇒ E =M v C2 + J ∆ c ω 2 ,
i =1 i =1 2 2
where J ∆ C is the moment of inertia of the rigid about an axis passing through the center of
mass, perpendicular to the fixed plane, and M is the rigid mass. The fact that
n n
∑m i x i = ∑m i y i = 0 has been taken into account, this sums representing the static
i =1 i =1
moment of the rigid about the planes Cxz and Cyz (see the theorem of the static moments).

15.3.4. Work-kinetic energy theorem in the motion of a system of material


points or rigid about a fixed reference

Statement (system of material points case): Variation in kinetic energy of a


system of material points in an infinitesimal time interval is equal to the sum of elementary
mechanical work of the external forces and elementary mechanical work of the internal
forces performed in the same time interval:

dE = dLext + dLint (15.46)

Proof: Let us consider the system of material points studied in paragraph 14.1.2.
Performing the scalar multiplication of the relations (15.7) by the differentials of the

position vectors d r 1 i , i = 1, n , and adding the relations thus obtained it is found that:
n → → n → → n n → →
∑ m i a i ⋅ d r 1i = ∑ F i ⋅ d r 1i + ∑ ∑ F i j ⋅ d r 1i (15.47)
i =1 i =1 i =1 j =1
j ≠i
The two terms in the right member of the relation (15.47) represent the work of the
external forces and the work of the internal forces, respectively, acting on the system:
n → → n n → →
dLext = ∑ F i ⋅ d r 1i , dLint = ∑ ∑ F i j ⋅ d r 1i (15. 48)
i =1 i =1 j =1
j ≠i
In addition:
→ →
n → → ndvi → n → d r 1i n → →
∑ m i a i ⋅ d r 1i = ∑m i
dt
⋅ d r 1i = ∑ m i d v i ⋅
dt
= ∑m i v i d v i =
i =1 i =1 i =1 i =1
n 1 →  n 1 
= ∑ d  m i v i2  = d  ∑ m i v i2  = dE
i =1  2   i =1 2 

248
Theoretical mechanics – theory and applications

From (15.46 – 15.48) we obtain (15.46).

By integration of the relation (15.46) between two times t 0 and t 1 we obtain the
work-energy theorem in total form

E 1− E 0 = L 0ext−1 + L int
0−1 (15.50)

In general, the elementary mechanical work of the internal forces is not null
although the internal forces are in pair and directly opposite. Considering the pair of
→ →
internal forces F i j and F j i , i, j = 1, n, i ≠ j , the corresponding mechanical work is:
→ → → → →  → →  → → →  → →
dLint = F i j ⋅ d r 1i + F j i ⋅ d r 1 j = F i j ⋅  d r 1i − d r 1 j  = Fi j ⋅  v i − v j  dt = F i j ⋅ v i j dt
   

→ → → d r 1i →
as F i j = − F ji and v i = . It was denoted by v i j the relative velocity of the point
dt
Ai related to point A j .
→ →
As dL int = 0 it is required that F i j ⋅ v i j = 0 . This happens if:

→ →
1) F i j = 0 - there is no interaction between Ai and A j ;
→ →
2) v i j = 0 - point Ai has the same velocity as the point A j (the case of two
bodies rolling one over the other without slippage);
→ →
3) F i j ⊥ v i j - internal force between points Ai and A j is perpendicular to their
relative velocity (case in which the distance Ai A j remains constant during the
motion, and point A j describes a motion on a sphere centered at Ai . Velocity

v i j will be perpendicular to sphere radius, namely on the line Ai A j which is

the line of action of the force F i j ).

In particular, if the system is a solid rigid (distance between any two points does
not change during the motion), the mechanical work of the internal forces is null.

It is thus obtained the following statement for the work-energy theorem:

Statement (rigid case): Variation in the kinetic energy of a rigid solid in an


infinitesimal time interval is equal to the elementary work of the external forces acting on
the rigid in the same time interval:

d E = dL ext (15.51)

249
Theoretical mechanics – theory and applications

As total form, the relation (15.51) is written as:

E 1− E 0 = L 0ext−1 (15.52)

15.3.5. Work-kinetic energy theorem in the motion of a system of material


points or rigid about the center of mass

Statement (system of material points case): Variation in kinetic energy of a


system of material points in its motion around its center of mass is equal to the work of
external forces and the work of internal forces applied to the system, computed with the
relative motions about the center of mass:

d E ' = dL 'ext + dL int


'
(15.53)
Proof: Koenig’s theorem for kinetic energy and the relation (15.20) are used (see also
paragraph 15.3.1). From (15.35) and (15.46) we find:

1 →2  n → → →  n n → → → 
d  M v C + E ' = ∑ F i ⋅ d  r C + r i'  + ∑ ∑ F i j ⋅ d  r C + r i'  ⇔
2  i =1   i =1 j =1  
j ≠i

 
→ → n →  → n → → n n →  →
M v C ⋅ d v C + dE ' =  ∑ F i  ⋅ d r C + ∑ F i ⋅ d r i' +  ∑ ∑ F i j  ⋅ d r C +
 i =1  i =1  i =1 j =1 
 j ≠i 
n n → →
+ ∑∑Fi j⋅d r '
i (15.54)
i =1 j =1
j ≠i
n n → → n → → not n n → → not
Dar ∑ ∑F i j = 0, ∑F i⋅ d r '
i = dL ext
'
, ∑∑Fi j⋅d r '
i = dL int
'
and
i =1 j =1 i =1 i =1 j =1
j ≠i j ≠i

→ → →d vC → → → → n →  →
M v C ⋅ d v C = M v C dt ⋅ = M d r C ⋅ a C = M a C ⋅ d r C = ∑ F i  ⋅ d r C
dt  i =1 
so that, from (15.54), we obtain (15.53)

'
Note: In the rigid case, dL int = 0 , so that relation (15.53) is reduced to

dE ' = dL 'ext (15.55)


which means that the work - kinetic energy theorem in the motion of a system of material
points or rigid about the centre of mass shall have the same form as with its motion about a
fixed point (same is valid as well for a system of material points).

250
Theoretical mechanics – theory and applications

15.3.6. Mechanical energy conservation

→ → → →
Given a force F = Fx i + F y j + Fz k , it is called conservative force if there is a
scalar function U : R 3 → R, U = U ( x, y, z ) so that:
∂U ∂U ∂U
Fx = , Fy = , Fz = (15.56)
∂x ∂y ∂z

Function U is called force function. Force F and its elementary work become:
→ ∂U → ∂U → ∂U →
F= i+ j+ k = grad U (15.57)
∂x ∂y ∂z
→ → ∂U ∂U ∂U
dL = F ⋅ d r = ⋅ dx + ⋅ dy + ⋅ dz = dU (15.58)
∂x ∂y ∂z
Let us consider a system of material points wherein each internal force
differentiates from a force function.
→ →
Thus, for the pair of internal forces F i j , F j i , i, j = 1, n, i ≠ j , there is the function
( )
U i j x1i , y1i , z1i , x1 j , y1 j , z1 j so that:
→ ∂U i j → ∂U i j → ∂U i j → → ∂U i j → ∂U i j → ∂U i j →
Fij = i+ j+ k , F ji = i+ j+ k
∂ x1i ∂ y1i ∂ z1i ∂ x1 j ∂ y1 j ∂ z1 j
The elementary work of the internal forces is:
 
n n → → n n n n 
dL int = ∑ ∑ F i j ⋅ d r 1i = ∑ ∑ dU i j = d  ∑ ∑U i j  = dU (15.59)
i =1 j =1 i =1 j =1  i =1 j =1 
j ≠i j ≠i  j ≠i 
n n
where U = ∑ ∑U i j is the force function of the system which depends on the position of
i =1 j =1
j ≠i
the points composing the system.
We define the potential energy of the system by the relation:
def
V = −U (15.60)
so that:
dL int = − dV (15.61)
Relation (15.46) (the mathematical expression of the work-kinetic energy theorem)
becomes:
dE = dL ext − dV or d ( E + V ) = dL ext (15.62)
Sum
def
E + V = E mec (15.63)
is called mechanical energy of the system.
If dLext = 0 , we obtain:

251
Theoretical mechanics – theory and applications

E mec = E + V = constant (15.64)


Relation (15.64) represents the theorem of mechanical energy conservation. Its
statement is:
If the elementary work of the external forces acting on a conservative system is
null for an interval of time, then the mechanical energy of the system is constant for that
interval of time.
15.4. Solved problems

R 15.1) A cylinder of the dimensions in Figure R 15.1.1 and weight G = 5000 N is placed
on a horizontal plane. Determine the work needed to roll over the cylinder around point A
of intersection of a generatrix with the horizontal plane.

1.2m C

C B B

O
1.6m

O 2
 
G G h
h1 D 2

D A
A
Figure R 15.1.1 Figure R 15.1.2

Solution: To roll the cylinder over it is necessary to bring its diagonal AC upright
(Figure R 15.1.2). The center of gravity rises from height h 1= 0,8 m to
height h 2 = h 12 + R 2 = 0,8 2 + 0,6 2 = 1 m . Work needed for roll over shall
be: L = G (h 2 − h 1 ) = 5000 ( 0,8 − 1) = −1000 J

R 15.2) A right circular cone of weight G, height h and radius r = h / 3 rolls without
slipping on a horizontal plane around its fixed vertex (Figure R 15.2.1). Determine the
kinetic energy of the cone if it rotates about the vertical axis passing through O at the
constant angular velocity ω = 4 π rad / s .

z y
h
r 
ω 

 ω0 O 2α G
ω α
2α 
O G 
ω1
x
Figure R 15.2.1 Figure R 15.2.2

252
Theoretical mechanics – theory and applications

Solution: The motion of the solid with fixed point is studied in relation to the
origin of a Cartesian system Oxyz with the origin O at the apex of the cone and the axis Oy
along the axis of symmetry of the cone (Figure R 15.2.2). The kinetic energy is given by
the relation:
1
( )
E = J x ω 2x + J y ω 2y + J z ω 2z − 2 J x y ω x ω y − 2 J y z ω y ω z − 2 J z x ω z ω x
2
(1)
where ω x , ω y , ω z are the projections of the angular velocity vector on the axes of the
system Oxyz. The absolute angular velocity results from a composition of concurrent
→ → → ω
rotations, ω a = ω + ω 1 . From the triangle of the angular velocities we obtain ω a = ,
tg α
so that the projections of the absolute angular velocity on the axes of the Cartesian
reference Oxyz will be:
ω x = 0 , ω y = −ω a cos α , ω z = ω a sin α (2)
It can be shown (see problems R 14.1 – R 14.3) that the axial and centrifugal
moments of inertia for the right circular cone related to the Cartesian system Oxyz are:
3 G  2 r 2  3 G 2
Jx = Jz =  h +  , Jy = r , J xy = J yz = J zx = 0 (3)
5 g 4  10 g
Entering (2) and (3) in (1) we find that:
3G 2  3 
E= ω cos 2 α  r 2 ctg 2 α + 6 h 2 + r 2  (4)
2 g  2 
h2 9 1
but ω = 4 π , cos 2 α = = , sin 2 α = , hence:
h +r
2 2 10 10
387 2 G 2
E= π h (5)
25 g
→ →
R 15.3) The weights P 1 and P 2 are connected by an inextensible wire of negligible
→ →
weight, passed over the fixed pulleys B and D. When the weight P 2 descends, weight P 1

ascends on the side AB of a prism ABDE of weight P 3 (Figure R 15.3). The angle made
by AB with the horizontal is α . Knowing that initially the system of the three bodies is at
rest and neglecting the frictions, find the travel of prism related to the floor for a travel h of

the weight P 2 . B C

C′3 C3
C′2
h C′1
x h
 C2
 C1 P3
P1 
A D
P2

Figure R 15.3

253
Theoretical mechanics – theory and applications

Solution: The system of bodies is acted upon by forces of gravity (vertical) only.
Center of mass does not move vertically. Noting by C 1, C 2 , C 3 the initial positions of the
centers of mass of the three bodies and by C 1' , C 2 ' , C 3 ' , respectively the positions of the

same points after the motion of the weight P 2 vertically by distance h, the abscissa x C of
the center of mass of the system corresponding to its two positions is determined. It is
considered that the prism moves to the left by the distance x.
P 1 x 1+ P 2 x 2 + P 3 x 3
At the initial instant: ( x C ) 0 = ;
P 1+ P 2 + P 3
P 1 ( x 1+ h cos α − x) + P 2 ( x 2 − x) + P 3 ( x 3 − x)
At the final instant: ( x C ) 1= .
P 1+ P 2 + P 3
P 1 cos α
As ( x C ) 0 = ( x C ) 1 we obtain: x = h.
P 1+ P 2 + P 3


R 15.4) A homogenous plate of weight G , bordered by a right triangle ABC of legs AB =
a and BC = b, rotates around a fixed axis containing leg BC (Fig. R15.4.1). At the initial
instant the angular velocity of the plate is ω 0 . Each element of the plate faces an air
resistance proportional to the area of the element and its speed while moving; the direction
of the resistance force is perpendicular to the element surface. Given the proportionality
factor k, determine the law of variation ω = ω ( t ) of the angular velocity, in the rotation
motion of the plate.
O1

R1
O1
a a
A B
B A

ω0 x dx
y

G  
b ω G 
b • dR

 C
R2
O2 O2

Figure R 15.4.1 Figure R 15.4.2


→ →
Solution: Denoting by R 1 and R 2 the reactions at the joints O 1 and O 2 and

by d R the air resistance force acting normally on the rectangular element of dimensions
dx and y, respectively, and applying the angular momentum theorem about the point B we
obtain :

d K B → →  → →  → → →  →
= M B R 1  + M B R 2  + M B R  + ∫A M B  d R  (1)
dt        

254
Theoretical mechanics – theory and applications

where the integral extends over the entire domain occupied by the plate (Figure R 15.4.2).
Projecting the vector equation (1) on the axis BC and noting that
→ →  → →  →
M BC  R 1  = M BC  R 2  = 0 (as the axis Bc contains points O 1 and O 2 ) and that
   
→ → →
M BC  G  = 0 (as BC is parallel to the weight direction), it results:
 
d K BC
= M R = ∫D x dR (2)
dt
The angular momentum of the plate ABC, in rotation motion, is determined using
the formula:
K BC = J BC ⋅ ω (3)
Ga2
It can be shown (see chapter13) that J BC = . The elementary force of
6g
resistance of the air
corresponding to an element of area dA
b
is dR = k ⋅ dA ⋅ v = k y dx v = k ( a − x) ω x dx , thus the resulting moment of the resistant
a
forces about the axis of rotation will be:
b a k a3
M R = − k ω ∫ a − x x dx = −
( ) ω (4)
a 0 6b
From (2-4) a differential equation with separable variables is obtained:
G dω a
= −k ω (5)
g dt b
Its particular solution (under initial conditions t 0 = 0, ω = ω 0 ) is:
 gka 
ω = ω 0 ⋅ exp −t (6)
 Gb 
R 15.5) A slider of weight G = 20 N attached to the fixed point O by means of a spring of
elastic constant k = 3 N/m is moving (downwards) along an upright straight bar (Figure R
15.5.1). Knowing that at the initial time the slide is at rest at point A and the length of the
spring in natural condition is l 0 = 4 m , determine the speed of the slider at the point B.
The distances AB = h = 6 m, OA = l 1= 8 m are also known.
l1
l1 A
A O

O k VA = 0 x
 θ
M
G 
 h
h Fel
  N
dr G
B
VB = ?
B
Figure R 15.5.1 Figure R 15.5.2

255
Theoretical mechanics – theory and applications

Solution: At instant t the slide is at the point M and has the speed v. The work-
kinetic energy theorem applies between the instants at which the slide is at points A and B:
E B − E A = L A− B (1)
1G 2
where E A = 0, E B = v (Figure R 15.5.2).
2 g B

The normal reaction is perpendicular to the displacement d r and does not
produce mechanical work, the mechanical work of the weight is L GA− B = G ⋅ h = 120 N m
and the work of the elastic forces is given by the relation:
→ →
L A−elB = ∫AB F el ⋅ d r = − ∫AB F el dr cos θ
F
(2)
But
cos θ =
AM
OM
=
x
, dr = dx, F el = k ∆ l = k l 12 + x 2 − l 0 ( )
l1+ x
2 2

so that:
F
h
L A−elB = − ∫ k (l 1+
2
x2 −l 0 )⋅ x  k x2
dx =  −
 2

+ k l 0 l 12 + x 2 

h
0
= −12 N m (3)
0 l 12 + x 2
We obtain v B = 10,29 m / s .
R 15.6) Let us consider the lifting mechanism in Figure R 15.6.1 consisting of
homogeneous bodies, connected by perfectly flexible and inextensible wires. Frictions are
19
negligible. To lift the load Q = G , the mechanism is actuated by a couple of
2
87
moment M = G R . Assuming that the system starts from rest, determine the law of
4
motion of the load Q. 
2R ,4G 
4  R ,4G
R, G
3
 O 4 = O5
M O3


R ,3G
2

O2

1

Q
x=?

Figure R 15.6.1

256
Theoretical mechanics – theory and applications

Solution: We shall use, on a case by case basis, the theorem of motion of the
center of mass and/or the theorem of the angular momentum.

Body of weight Q (Figure R 15.6.2)

The theorem of the motion of the center of mass projected on Oy:


Q
a = S 1− G (1)
g
S1 a =a S
1 2 S3

O2

Q S1
ε2
3G
Figure R 15.6.2 Figure R 15.6.3

Body of weight 3 G (Figure R 15.6.3)

The theorem of the motion of the center of mass projected on Oy:


3G
a = S 2 + S 3 − S 1− 3G (2)
g 2
The angular momentum theorem about point O 2 , projected on O 2 z :
J O 2 ε 2= − S 3 R + S 2 R (3)

Body of weight 4G (Figure R 15.6.4)

The angular momentum theorem about point O 3 , projected on O 3 z :


J O 3 ε 3= S 4 R − S 3 R (4)

M0 S4
V4
V3
H3 O4
S4 O3 H4 S2
4G
ε3 5G
S3 ε4

Figure R 15.6.4 Figure R 15.6.5

257
Theoretical mechanics – theory and applications

Body of weight 5G (Figure R 15.6.5)


The angular momentum theorem about point O 4 , projected on O 4 z :
J O 4 ε 4 = M 0 − S 4⋅ 2R − S 2 R (5)
But:
3G R 2 2G R 2 4G ⋅ ( 2 R) 2 G R 2 17G R 2
JO 2 = , J O3= , J O4= + = (6)
2g g 2g 2g 2g
and the kinematic study leads to the relations:
a 4a 2a
a 2= a , ε 2= , ε 3= , ε 4= (7)
3R 3R 3R
From (1-7) the weight acceleration Q is obtained:
69 G − 6 Q
a= g (8)
41 G + 6 Q

15.5. Problems proposed

15.5.1. Conventional tests


→ →
TC 15.1) Two boats of weight G are moving in the same sense at the same velocity v .

At an instant, from the first boat a weight P is launched towards the second boat, at the

velocity u (relative to the boats). Determine the velocities of the two boats upon launch of
the weight and receiving of the weight, respectively.

TC 15.2) A wheel of radius r which rotates with constant angular velocity ω 0 around its
axis of symmetry is pressed by a brake shoe AB with the constant radial force F (Figure
TC 15.2.1). Knowing that the moment of inertia of the wheel about the axis of rotation is J
and the wheel stops after t 1 seconds as a result of friction between it and the shoe, it is
required:
a) The coefficient of friction µ between shoe and wheel;
b) Number of wheel revolutions performed until stopped.

ω

A
r   x
F Mg
O 
ω0 k
B
 B
mg

Figure TC 15.2.1 Figure TC 15.3.1

258
Theoretical mechanics – theory and applications

TC 15.3) An inextensible thread bearing at one end a load of mass m passes over a pulley
which rotates around the horizontal axis Oz. The other end of the thread is attached to a
vertical spring having extremity B fixed. The tension force of the spring is proportional to
its elongation, the proportionality factor being k (Figure TC 15.3.1). Determine the load
oscillation period knowing that the pulley mass is M and that the thread does not slip on
the pulley.

TC 15.4) Let us consider the system of objects in Figure TC 15.4.1, starting from rest
under the action of its own weights. The disc of weight Q and radius R can move on a
horizontal plane and is connected through a flexible and inextensible thread to a body of
weight P, located on a plane inclined at the angle α with the horizontal. The thread is
winding on the disk of radius r of a winch and unwinds from the disk of radius R of the
same winch. The moment of inertia of the winch about the axis of rotation is J.
Considering that the disc of weight Q rolls without slipping on the horizontal plane, the
coefficient of rolling friction being s and the body of weight P moves on the inclined plane
with friction, the coefficient of sliding friction being µ , study the motion of weight P and
determine its acceleration.
 2r
3
r, R , J z , G
2 r, R , J
C
R
O1 O

O2 B
Q 1
2b
a =?
I(G )
 h
Q
α

P

Figure TC 15.4.1 Figure TC 15.5.1

TC 15.5) A homogenous plate of weight P, width 2b and height h, can rotate around the
vertical axis AB, on which there is also the cylinder C of radius r and negligible weight
(Figure TC 15.5.1). On this cylinder a wire is wound whose end passes over a pulley of
radius r of a winch having the moment of inertia J z 1 about the axis of rotation. The
system is driven by the weight Q, hanging at the end of the thread passing over the pulley
of radius R of the winch. During the motion each element of the area of the plate
undergoes the resistance of the air, which is proportional to the area and speed of the
element, the proportionality factor being k.
a) Study the motion of the system and determine the law of motion and the law of variation
of the angular velocity of the plate;
b) Determine the tension in the threads;
c) Assuming that at an instant t 1 the weight Q pulls off, determine the time t 2 after which
the angular velocity is halved.

259
Theoretical mechanics – theory and applications

15.5.2. Multiple-choice tests



TG 15.1) A boat of weight G and length l is at rest and touches the pier with the bow

(Figure TG 15.1). A man of weight P at that instant in the middle of the boat starts to
move towards the shore. Determine the distance the boat will depart from shore when the
man reaches the end of the boat.
P l P l G l P l
a) x = ⋅ ; b) x = ⋅ ; c) x = ⋅ ; d) x = ⋅ .
G 2 P+G 3 P+G 2 P+G 2

l /2 l /2 x


 P
P
 
G G

Figure TG 15.1

TG 15.2) A pendulum is allowed to swing freely in the vertical plane, from the initial
position given by the angle θ 0 (Figure TG 15.2). Knowing the material point mass m and
length l of the wire, determine the period T for the small oscillations.
m π l l l
a) T = 2 π ; b) T = ; c) T = 2 π ; d) T = π .
g 2 g g 3g

θ0
θ
l
•m

Figure TG 15.2

15.6. Indications and answers

TC 15.1) Since the forces acting on the system are vertical, the linear momentum in
direction x remains constant (Figure TC 15.1). Conservation of momentum is applied to
the system consisting of boat 1 and weight 3 between instants t 1 and t 2 , respectively, when
the weight 3 is in the boat and in the air, respectively:

260
Theoretical mechanics – theory and applications

v + v = v 1+ (v − u )
G P G P
H 1= H 2 ⇒
g g g g
P
It is obtained: v 1= v + u>v.
G
To determine the speed v 2 of the boat 2 after receiving the weight, the momentum
conservation is applied for the system consisting of boat 2 and weight 3 between instants
t 2 and t 3 , respectively, when the weight is in the air and in the boat 2, respectively:

H '2 = H '3 ⇒ (v − u ) + v = v 2 + v 2
P G P G
g g g g
P
It results: v 2 = v − u<v .
P+G
y

u 3

• 
A

x
r T 
 F
P
O  
N −N
 
 
ω −T B
3
C1 V C2 V
1

  R
G G
 
V2 V1

Figure TC 15.1 Figure TC 15.2.2

TC 15.2) a) By singling out the two bodies (shoe and wheel) we obtain the situation in
Figure TC 15.2.2. The equations of equilibrium for the shoe are:

∑ X i =N − F = 0 , ∑Y i = R − T = 0 , T =µ N (1)

By applying the angular momentum theory about point O and projecting the
obtained vector relation on the direction of the axis of rotation we find the equation of
motion for the wheel:
Jε = −T r (2)
Integrating equation (2) with respect to time and considering the initial
condition t = 0 : ω = ω 0 , the law of variation of the angular velocity is obtained:
µrF
ω (t ) = ω 0 − t (3)
J
The coefficient of friction µ is determined by imposing the condition that the
wheel stops after t 1 seconds: t = t 1: ω = 0 . It results:
Jω 0
µ= (4)
r t1 F

261
Theoretical mechanics – theory and applications


b) Noting that ω = and integrating again the equation (3) with respect to time, the law
dt
of motion of the wheel is obtained:
µ rF 2
φ (t ) = ω 0 t − t (5)
2J
The number of wheel revolutions performed until stop is determined by noting that
a complete revolution corresponds to an angle φ = 2 π . The number is:
φ (t 1 ) ω 0 t 1
n rotatii = = (6)
2π 4π

TC 15.3) Denote by x the travel of the body of mass m at the instant t (Figure TC 15.3.2).

(1) Body 1: m a = m g − T (center of mass motion theorem)


(2) Body 2: J O ε = T R − Fel R (angular momentum theorem about O)

ε = a/R R0

T

2
O
1

a = x x
 
 −T Fel
mg 
Mg

Figure TC 15.3.2

MR2 ••
But J O = , F el = k x , a = x , so that from (1) and (2) we obtain the differential
2
equation:
•• 2k 2m
x+ x= g (3)
M + 2m M + 2m

2k
Denoting by p = the own pulsation of oscillation, the motion period will
M + 2m
be:
2π M + 2m
T= = 2π (4)
p 2k

TC 15.4) The sense of motion of the system is given by the weight P, going down the
incline. To study the motion, the total work-kinetic energy theorem is applied:

262
Theoretical mechanics – theory and applications

E f − E i = L i, f (1)
where :
Ei = 0 - the kinetic energy of the system at the start of the motion. It is zero
because the system starts from rest.
Ef - the kinetic energy of the system at an arbitrary instant of time t. The
weight P has travelled the space x at that time.
L i, f - the work of external and constraint forces acting on the bodies of the
system between the two time instants.
The connection between the kinematic characteristics of the motion of bodies is
given in the table below (see also Figure TC 15.4.2) .
θ3 , ω3 , ε3 θ 2 , ω2 , ε 2

v3 , a 3
O3

 RO 
 Q
2
N1
T3
O2

Mr 
N3 v1 , a1
α

P

Figure TC 15.4.2

Body Motion type Distance Velocity Acceleration


1 Translation x 1= x • ••
v 1= x a 1= x
2 Rotation x • ••
θ 2= x x
R ω 2= ε 2=
R R
3 Plane – r
x 3= r • r ••
parallel x v 3= x a 3= x
R R R
θ 3=
r r • r ••
x ω 3= x ε= x
R2 R2 R2

The kinetic energy of the three bodies at instant t is:


1 1P•2
E 1= m 1 v 12 = x
2 2g
2
•
1 2 1 x
E 2= JO 2 ω = J   (2)
2 2 R
 

263
Theoretical mechanics – theory and applications

2 2
1 2 1 2 1 Q  r •  1 Q R 2  r • 
E 3= m 3 v 3 + J O 1 ω 3 = x + x
2 2 2 g  R  2 g 2  R 2 

The kinetic energy of the system at the same instant is obtained by summating the
kinetic energy of the component bodies:
1 P 3 r2 J  • 2
E f = E 1+ E 2 + E 3=  + Q+ x (3)
2  g 2g R 2 R 2 

The only forces giving mechanical work are the force of weight P and the friction
force T 1 , the other forces being normal to the displacement ( Q , N 1 , N 2 ), with fixed point
of application ( RO 2 ) or applied at the instantaneous center of rotation ( T 3 ). The nonzero

work is also obtained due to the rolling friction moment M r . Summating these mechanical
works we obtain:
L i , f = P (sin α − µ cosα ) x − M r θ 3 (4)
where M r = s N 3= s Q .
From relation (4) and the table above we obtain:
 r 
L i − f =  P (sin α − µ cosα ) − s Q x (5)
 2
 R 
denoting:
P 3 r2 J r
A= + Q+ , B = P (sin α − µ cos α ) − s Q (6)
g 2g R2 R2 R2
relation (1) becomes :
1 •2
A x −0 = Bx (7)
2

Differentiating this differential equation with respect to time and simplifying by x , we
find the value of the acceleration of the body of weight P:

P (sin α − µ cosα ) − s Q
r
••
a 1= x =
B
= R2 (8)
A P 3Q r2 J
+ +
g 2g R 2 R2
Note: From the relation above and table T 12.8 one may conclude that all of the bodies
of the system have uniformly accelerated motions (of constant accelerations).

TC 15.5) a) Single out the three bodies of the system, introduce the given external forces
and constraint forces and apply theorems of the center of mass motion and angular
momentum (depending on the motion of the body under study).

264
Theoretical mechanics – theory and applications

Body of weight Q (Figure TC 15.5.2)

The theorem of center of mass motion projected on Oy:


Q
a =Q−S1 (1)
g 
y V0 
 ε2
S2
S1 
H0
a1 = a ⋅ O2
z

− S1 

Q G

Figure TC 15.5.2 Figure TC 15.5.3

Winch (Figure TC 15.5.3)

The theorem of angular momentum about O 2 on O 2 z :

J z 1 ε 2 = S 1 R − S2 r (2)

Flat plate and cylinder (Figure TC 15.5.4)

z

− S2 2r
ε3 
R 2y

R 2x y
dy
h O3 dR = kvdA

 y
x P
dA = hdy
2b 
R1z
 
R1x R 1y

Figure TC 15.5.4

The theorem of angular momentum about O 3 on O 3 z :

J z 2 ε 3= S 2 R − M R (3)

265
Theoretical mechanics – theory and applications

where M R is the moment due to air resistance forces and


a 2
ε 3= ε 2 =
R A ∫ A A ∫ 3 ∫
, M R = y ⋅ k v⋅ dA , J z 3 = y 2 dm = y 2⋅ ρ h dy = ρ h b 3 (4)

2 P 2
But P = M plate ⋅ g = 2 ρ h b g , so that J z 2 = b .
3 g
Moment M R is computed as follows:
b b 2 dy = 2 k h b 3 ω
M R= k ∫− b y ⋅ ω y ⋅ h dy = h k ω ∫− b y 3
(5)
as v = ω y , dA = h dy .
not
Denoting ε 2 = ε 3 = ε , from equations (1 - 3) we obtain the values of the tension in
the threads as a function of the angular acceleration ε :
 εR  εRR ε
S 1= Q 1 −  , S 2 = Q 1 −  − J z 1 (6)
 g   g r R
 •
as well as the differential equation with the angular velocity ω  ε = ω  as unknown:
 
 
•.
α ω+ β ω = Q R (7)


, β = h k b 2 . Since ω = ϕ , where ϕ = ϕ (t ) is the
Q 2 2P 2 2
where α = R + Jz1 + b
g 3g 3
angle made by the plate plane with its former position at the initial instant, we obtain the
following non homogenous second order linear differential equation:
•• •.
α ϕ + βϕ =QR (8)

The general solution of this equation is:

  −βt  
Q R α  α 
ϕ (t ) = e + 1 + t  (9)
β β   
   
and represents the law of motion of the flat plate. The angular velocity and the angular
acceleration in the rotation motion of the plate are obtained by differentiating the law of
motion (9):
 β β
. QR − t  . QR −α t
ω (t ) = ϕ (t ) = 1 − e α
 , ε (t ) = ω (t ) = e (10)
β   α
 

266
Theoretical mechanics – theory and applications

b) The tensions in threads result from the relations (6):

 β β
 Q R2 −α t  
R   Q R 2 J z 1  − α t
S 1= Q 1 − 
e  , S 2 = Q 1 −  +

e

(11)
 g α  r α g α
     

2P 2
c) For Q = 0, the value of the constant α becomes α 1= J z 1 + b and the differential
3g
•• •
equation becomes homogenous, α 1 ϕ + β ϕ = 0 , and has the general solution:
β
− t
ϕ (t ) = C + C e α 1 (12)
1 2
.
Imposing the initial conditions, t = 0 : ϕ = 0 , ϕ = ω 1 , we obtain the particular solution
 β 
 − t
α1
ϕ (t ) = ω 1 1 − e α 1  . Instant t 2 is determined requesting that the angular velocity
β  
 
.
ω =ϕ is half of the value ω 1 recorded at the instant of weight Q pull off,
β
− t
. ω
namely ϕ (t 2 ) = ω 1 e α 1 = 1 , thus:
2
α1
t 2= ln 2 (13)
β

TG 15.1) The forces acting on the bar-man system are vertical so that the momentum (and
hence velocity of the center of mass) is conserved horizontally. Since at the initial instant
the system is at rest the velocity of the center of mass remains zero throughout the motion,
i.e. the center of mass does not change during the motion:

P Q l
x + x + 
g 2
( x C )t =t initial = ( x C )t =t final l g
⇒ = ⇒ x=
P l
⋅ .
2 P Q P+G 2
+
g g
Correct answer: d)

TG 15.2) Let θ be the angle formed by the thread with the vertical direction at the arbitrary
instant t. The angular momentum theorem is applied about the thread attachment point O:
J O ⋅ ε = ∑ M O (*). The only force which gives a moment about O is the weight
•• g
G = mg and J O = m l 2 . From (*) we find the differential equation θ + sin θ = 0 . For the
l

267
Theoretical mechanics – theory and applications

small oscillations of the thread sin θ ≅ θ may be considered, hence the equation of
•• g
motion θ + θ = 0 . Its solution is of the form θ ( t ) = C 1sin p t + C 2 cos pt ,
l
g
where p = is the pulsation of the motion.
l

As T = , the correct answer is c).
p

268
Theoretical mechanics – theory and applications

16. Dynamics of the rigid


16.1. Dynamics of rigid with fixed axis

16.1.1. Problem formulation

Let us consider a rigid solid (C) having two points, O 1 and O 2 , fixed (i.e.
line O 1 O 2 is unmovable. Suppose that the rigid is acted upon by a system of arbitrary
→ → → 
external forces F i , i = 1, n , which can be replaced by the torsor  R , M O 1  . Under the
 
influence of these forces the rigid will develop a motion of rotation around the axis
∆ = O 1O 2 (Figure T 16.1).

z1 , z

R2 O2

zc MO
1 
 R
R1 d
C
y
θ
O1 = O y1
x1 xc
θ

x
(Δ )

Figure T 16.1
Knowing the rigid motion means knowing the law θ = θ ( t ) ,
 ∧  → →
where θ = masO 1 x 1, Ox  . In addition, of importance will be the reactions R 1 and R 2 at
 
the fixed points (spherical joints) O 1 and O 2 .
The study of motion will be made with respect to the triorthogonal trihedrals
O 1 x 1 y 1 z 1 (fixed) and Oxyz (movable, stiff-connected with the rigid) having O 1≡ O , axes
O 1 z 1≡ O z ≡ (axis of rotation ∆ ) and the center of mass C of the rigid contained in the
movable plane Oxz.

Denoting by R 1x , R1 y , R1z and R 2 x , R2 y , R2 z the projections of the reactions R 1

and R 2 on the axes of the movable reference, the unknowns of the problem are in number
of seven (7), namely: θ , R 1x , R1 y , R1z , R 2 x , R2 y , R2 z .

269
Theoretical mechanics – theory and applications

16.1.2. Study of rigid motion


The work-kinetic energy theorem is used about the fixed point O 1 . The constraint
→ → →
forces R 1 and R 2 , respectively and the resultant R of the external forces do not produce
mechanical work because the points O 1 and O 2 are fixed.

The elementary mechanical work produced by M O 1 is equal to:
→ → → →
dL ext = M O 1 ⋅ d θ = M O 1 ⋅ ω dt = M z ω dt (16.1)
The rigid has a rotation motion so that the kinetic energy is expressed as:
1 1
E = J ∆ ω 2= J z ω 2 (16.2)
2 2
From (16.1), (16.2) and (15.51) we find:
1 
d  J z ω 2  = M z ω dt
2 
or
••
J z ε =M z ⇔ J zθ =M z (16.3)
1  1 dω
as d  J z ω 2  = J z ⋅ 2 ω dω = J z ω ⋅ dt = J z ω ε dt .
2  2 dt
 • 
Integration of the differential equation (16.3), where M z = M z θ , θ , t  , allows
 
obtaining of the law of motion θ = θ ( t ) .

16.1.3. Determination of reactions



(
In order to obtain the reactions R 1 R 1x , R1 y , R1z ) →
(
and R 2 R 2 x , R2 y , R2 z ) the
linear momentum and angular momentum theorems are applied.
Theorem of the (linear) momentum: Denoting by R x , R y , R z the projections of the

resultant R of the external forces on the axes of the movable reference Oxyz, the theorem
of the momentum (relation 15.5) becomes:

→ → → →
H = R + R 1+ R 2 (16.4)
or, in terms of projections on the axes of the movable reference Oxyz,
• • •
H x = R x + R 1x + R2 x , H y = R y + R 1 y + R2 y , H z = R z + R 1z + R2 z (16.5)
But
→ → →
i j k
→ → → → →
H =M vC =Mω× rC =M⋅ 0 0 ω = M ω xC j
xC 0 zC

270
Theoretical mechanics – theory and applications

So that:
→ → →
• → → i j k
→ d H ∂H → → →
H= = + ω × H = M ε xC j + 0 0 0 ⇒
dt ∂t
0 M ω xC 0

→ → →
H = − M ω 2 xC i + M ε xC j (16.6)
From (16.5) and (16.6) we obtain the system:
− M ω 2 x = R + R + R
 C x 1x 2x
 M ε xC = R y + R 1 y + R2 y (16.7)

0 = R z + R1 z + R2 z

Theorem of the angular momentum about the fixed point O 1 : From (15.28) we
find that:

→ →→ →  → → 
K O1 = M + M R 1 + M O1R 2  (16.8)
   
O1 O 1

Denoting by M x , M y , M z the projections of the resultant moment of the


→ →  →
external forces about the point O 1 , noting that M O 1  R 1  = 0 and that:
 
→ → →
i j k
→ →  → → → →
M O 1  R 2  = O 1O 2 × R 2 = 0 0 d = − d R 2 y i + d R2 x j (16.9)
 
R 2x R 2y R 2z

where d = O 1O 2 , from (16.8) we obtain by projections on the axes of the reference Oxyz
the equations:
• • •
K x= M x− d R 2 y , K y = M y + d R2 x , K z= M z (16.10)
But (see relation 15.23)
→ → → → → → →
K O1= K x i + K y j + K z k = −J xz ω i − J y z ω j + J z ω k
so that:
• →
→ ∂K O1 → → → → →
K O1 = + ω × K O1 = −J x z ε i − J y z ε j + J z ε k + (16.11)
∂t
→ → →
i j k
+ 0 0 ( ) (

ω = − J x z ε + J y zω 2 i + − J y z ε − J x zω 2 j + J z ε k ) → →

−J xz ω − J yz ω J z ω

271
Theoretical mechanics – theory and applications

From (16.10) and (16.11) we obtain the system:


− J ε + J ω 2 = M − d R
xz yz x 2y

− J y z ε − J x z ω 2 = M y + d R 2 x (16.12)

 J z ε = M z

Note: The last equation of the system (16.12) coincides with the equation of
→ →
motion (16.3). To determine the components of the reactions R 1 and R 2 we have
available only five equations (equations (15.8) and the first two equations (16.12)) hence a
component of the reactions ( R 1z or R 2 z ) remains undetermined. In practice, in order to
remove this indetermination the spherical joint in O 2 is replaced with a cylindrical one
( R 2 z = 0). From the last relation (16.7) it follows that R 1 z = − R z .

16.1.4. Rotors balancing

A rigid solid in a rotation motion is called rotor. In practice it is desirable that the
→ →
reactions R 1 and R 2 in the joints bearings are as small as possible in order that the wear
of materials from which they are made to be minimized. From equations (16.7) and (16.12)
it is observed that these reactions depend on the motion by factors ε and ω 2 , theoretically
being possible to take very high values. Naturally the question arises whether it is possible
to take actions so that the increase of the angular velocity vector module does not have as
result the increase of intensity of the components normal to the fixed axis, corresponding
to the two forces immobilizing the axis.
Let us consider the system of equations that corresponds to the rest condition of the rigid,
i.e. the system obtained from (16.7) and (16.12) for:
ω=ε =0
therefore also for M z = 0 :

0 = R x + R 1stx + R st
2x , 0 = R y + R 1sty + R st
2y , 0 = R z + R 1stz + R st
2z
(16.13)
0 = M x− d R 2sty , 0 = M y+ d R 2stx

The reactions corresponding to the rest condition have been denoted


by R 1stx , R 1sty , R 1stz , R 2stx , R 2sty , R st
2z and are called static reactions. The reactions
corresponding to the motion condition are called dynamic reactions and will be denoted
by R 1dx , R 1dy , R 1dz , R d2 x , R d2 y , R d
2z . They are determined from the system composed of the
equations (16.7) and (16.12).

272
Theoretical mechanics – theory and applications

In practice the interest is to make rotors whereto the dynamic reactions are equal to
the static ones, i.e. rotors for which the motion does not affect the intensity of reactions.
Such a rotor is called balanced.
The conditions for a rotor to be balanced are obtained by replacing the dynamic
reactions with the static ones, in the system that solves the problem of dynamics of rigid
with fixed axis:
− M ω 2 x = R + R st + R st − J ε + J ω 2 = M − d R st
 C x 1x 2x

xz yz x 2y

 M ε xC = R y + R 1 y + R 2 y
st st
, − J y z ε − J x z ω = M y + d R 2stx
2
(16.14)
 
0 = R z + R 1z + R 2 z
st st  J z ε = M z

From (16.13) and (16.14) it follows:

− M ω 2 x C = 0 , M ε xC = 0
(16.15)
− J xz ε + J y z ω = 0 , − J yz ε − J x z ω = 0 ,
2 2
Jz ε = M z
Since, in general, ω ≠ 0 and ε ≠ 0 , from the first two relations (16.15) we obtain
the first condition for rotor balancing:
xC = 0 (16.16)
This means that the axis of rotation should be the central axis, i.e. to contain the
center of mass C. A rotor for which the condition (16.16) is fulfilled is called statically
balanced.
The third and fourth relation (16.15) form a homogeneous linear system in the unknowns
ω 2 and ε . As ω ≠ 0 and ε ≠ 0 this system should also have nonzero solutions, condition
met if its determinant is zero:
− Jxz Jyz
=J xz+
2
J y z=
2
0 ⇔ J xz = J yz = 0 (16.17)
− Jyz − Jxz

From (16.17) we find that the rotation axis Oz shall have to be the main axis of
inertia. A rotor for which this condition is met is called dynamically balanced. In
conclusion, for a rotor to be balanced both statically and dynamically it is necessary and
sufficient that the axis of rotation is a central and principal axis.

16.2. Dynamics of the rigid with a fixed point

16.2.1. Problem formulation

Let us consider a rigid solid with O as fixed point, which is subject to the action of
→ → 
a system of arbitrary external force that reduces at O to the torsor τ O R, M O  .
 

The rigid has at a spherical joint at O that is replaced by the reaction R ' of unknown
magnitude and direction, according to the principle of action and reaction (Figure T 16.2).

273
Theoretical mechanics – theory and applications

The study of motion around the fixed point O is made in relation to two reference
systems, namely:
- the fixed system O 1 x 1 y 1 z 1 originating in O ( O ≡ O 1 );
- the movable system OXYZ, stiff-connected to the rigid, with Ox, Oy and Oz as principal
axes of inertia with respect to point O. The rigid position relative to the fixed reference is
determined by Euler angles ψ , ϕ and θ (see Kinematics, Chapter 8).
The rigid motion is known if the following functions are known:
ψ = ψ (t) , ϕ = ϕ (t) , θ = θ (t) (16.18)
Besides these unknowns it is also necessary to determine the components of the

( )
reaction R ' R 'x , R 'y , R 'z on the axes of the movable reference Oxyz, hence the unknowns
of the problem are in number of six.

z1 z1
z z

M0
θ
 y
R ϕ
y φ ψ
 N′
 
R′ R′
O1 = O y1 θ O1 = O y1

ψ
x1 x1 ϕ
x N x

Figure T 16.2 Figure T 16.3

16.2.2. Finding of the equations of motion. Relations between the instantaneous


angular velocity vector components and Euler angles

To study the motion the angular momentum theorem is used about the fixed point

O. Since the line of action of the reaction R ' contains point O, this force does not give a
moment about point O.
From (15.28) we obtain:

→ →
K O= M O (16.19)
We have considered that Ox, Oy and Oz are principal axes of inertia and therefore
the expression of the angular momentum is:
→ → → →
K O = J 1ω x i + J 2 ω y j + J 3ω z k
where J 1, J 2 , J 3 are the principal moments of inertia about the fixed point O.

K O is a vector related to the axes of the movable system, hence:

274
Theoretical mechanics – theory and applications

→ → →
• → i j k
→ ∂K O → → → → →
K O= + ω × K O = J 1ε i + J2 ε j + J3 ε z k + ω x ωy ωz =
∂t x y
J 1ω x J 2ω y J 3ω z

[
= J 1 ε x + (J 3− J 2 )ω y ω z ] i + [J 2 ε y + ( J 1− J 3 )ω x ω z ] j + [J 3ε z + ( J 2 − J 1 )ω y ω x ] k
→ → →

(16.20)

Denoting by M x , M y and M z the components of the resultant moment M O of
the external forces, from (16.19) and (16.20) we obtain the system of the differential
equations of motion:
J 1 ε x + ( J 3 − J 2 )ω z ω y = M x
J 2 ε y + ( J 1− J 3 )ω x ω z = M y (16.21)
J 3 ε z + ( J 2 − J 1 )ω y ω x = M z

known also as Euler equations for a rigid with a fixed point.


For integration of the equations (16.21) the relations of connection between the
angular velocity vector components ω x , ω y , ω z and Euler angles ψ , ϕ and θ are needed.
The angles ψ , ϕ and θ are independent, modification of any of them not affecting the
values of the other two. Variation of each angle corresponds to an independent rotation
• • •
with the angular velocityψ , ϕ ,θ , respectively (Figure T 16.3). A rotation around the axis
O 1 z 1 determines merely the change in the angle ψ , the corresponding angular velocity

vector having the direction of the axis O 1 z 1 and the scalarψ . A rotation around the axis
Oz leads only to change in the angle ϕ , the corresponding angular velocity vector being

oriented on Oz and having the scalar ϕ . Eventually, a rotation around the line of nodes ON
determines only the change in the angle θ . The corresponding angular velocity vector has

the direction ON and the scalar θ . Composing the angular velocity vectors corresponding

to the three rotations the angular velocity vector is ω obtained. Thus:
→ • → • → • →
ω = ψ ⋅ k 1+ ϕ ⋅ k + θ ⋅ u ON (16.22)

where u ON is the versor of the axis of nodes.
Line O 1 z 1 belongs to the plane (ON ‘z), where ON ' ⊥ON , so that:
→ → →
k 1 = cos θ ⋅ k + sin θ ⋅ u ON '
→ → →
Since u ON ' = sin ϕ ⋅ i + cos ϕ ⋅ j , we obtain:

→ → → →
k 1= sin θ sin ϕ ⋅ i + sin θ cos ϕ ⋅ j + cos θ ⋅ k (16.23)

275
Theoretical mechanics – theory and applications

In addition,
→ → →
u ON = cos ϕ ⋅ i − sin ϕ ⋅ j (16.24)

From (16.22) – (16.24) we obtain the components of the angular velocity ω on the
axes of the movable reference Oxyz:
 • •
ω
 x = ϕ sin θ sin ϕ + θ cos ϕ
 • •
ω y = ϕ sin θ cos ϕ − θ sin ϕ (16.25)
 • •
ω = ϕ + ψ cos θ
 z

Functions ψ = ψ ( t ) , ϕ = ϕ ( t ) , θ = θ ( t ) are obtained by solving the system


• • •
(16.21), with ω x , ω y , ω z given by (16.25) and ε x = ω x , ε y= ω y , ε z = ω z . Integration
of the system (16.21) could not be accomplished for any other set of initial conditions than
three particular cases, namely:
→ →
i) Euler – Poinsot Case: It is considered that M O = 0 (that is
M x = M y = M z = 0 ). Such a situation is obtained in the case of a rigid acted upon by its
own weight only and suspended from the center of mass.
ii) Lagrange – Poisson Case: It is considered that the center of mass is on the
movable axis Oz ( x C = y C = 0) and that J 1= J 2 . The rigid is acted upon by its own weight
only.
iii) Sofia Kovaleskaia Case: The center of mass is in the plane Oxy ( z C = 0) ,
J 1= J 2 = 2 J 3 and the rigid is acted upon by its own weight only.

16.2.3. Determination of the reaction



Reaction R ' is obtained by applying the linear momentum theorem:

→ → →
H = R+ R ' (16.26)
→ → → → 
But H = M v C = M  ω × r C  and
 
• →
→ ∂H → → → → → → → 
H= + ω × H = M ε × r C + M ω × ω × r C 
∂t  
So that, from (16.26), we find:

→ → → → → → → → → 
R ' = − R + H = − R + M ε × r C + M ω × ω × r C 
 

276
Theoretical mechanics – theory and applications

16.2.4. Gyroscope. Stability. Gyroscopic effect.


Regular precession motion. Gyroscopic moment.
16.2.4.1. Generals
Gyroscope is a body of revolution having a fixed point O, which rotates at a very
high angular velocity around its axis of symmetry (that contains point O). It is considered
in addition that the (principal) moment of inertia about the axis of symmetry, denoted J 3 ,
is much larger than the other two principal moments of inertia J 1 and J 2 (equal to each
other):
J 3 >> J 1= J 2 (16.28)
The Gyroscope can be:
a) Centered gyroscope – if the fixed point O coincides with the center of mass C
and the only force acting on the rigid is its own weight;
b) Not centered gyroscope– if the center of mass does not coincide with the fixed
point O but is on its axis of symmetry and the only force acting on the rigid is
its own weight.

16.2.4.2. Stability of the centered gyroscope


The centered gyroscope is a customization of the case Euler - Poinsot. Suspension
of the gyroscope is achieved by means of a system of gimbals that allows its simultaneous
rotation around three axes perpendicular to one another and concurrent at point O (which
coincides with the center of mass). The axis of symmetry is considered the Oz axis of the
movable trihedral stiff-connected to the gyroscope.
Since the moment of the only force acting on the rigid (its weight) about the fixed point O
is zero and J 1= J 2 , Euler's equations (16.21) become:
J 1 ε x + ( J 3 − J 1 )ω z ω y = 0 , J 1 ε y + ( J 1− J 3 ) ω x ω z = 0 , J 3 ε z = 0 (16.29)

From J 3 ε z = 0 we obtain ε z = ω z = 0 , that is:
not
ω z = ω 0 = constant (16.30)
From (16.30) and the second equation (16.29) we find that:
• J 3− J 1
ε y= ω y= ω0 ωz (16.31)
J1
Differentiating the first equation (16.29) and using (16.30) we obtain:
••
J 1 ω x + ( J 3− J 1 ) ω 0ω y = 0 (16.32)
Substituting (16.31) in (16.32) the following differential equation is obtained with
the unknown
••
ω x+ p 2 ω x= 0 (16.33)
J 3− J 1
We used the notation p = ω 0 . The general solution of this equation is:
J1
ω x = C 1 cos pt + C 2 sin pt (16.34)

277
Theoretical mechanics – theory and applications

where C 1 and C 2 are integration constants that depend on the initial conditions of the
motion. From the first equation (16.29) and (16.34) the component ω y of the instantaneous
angular velocity is also determined:

( J 3 − J 1 ) ω 0 (− C 1 p sin pt + C 2 p cos pt )
J1
ω y= −
or
ω y = ± (C 1sin pt − C 2 cos pt ) (16.35)
on how J 3 > J 1 or J 3 < J 1 .
If the gyroscope is not perturbed, then at the initial instant t 0 = 0 the conditions
are met as follows:
t 0 = 0 : ω x (t 0 ) = ω y (t 0 ) = 0 (16.36)
From (16.35)-(16.36) we find that, in these conditions:
ω x = ω y = 0 , ( ∀) t
→ → →
ω =ω z k =ω 0 k (16.37)
the gyroscope being not deviated in the direction Oz. Gyroscope motion is stable. If the
( ) ( )
gyroscope is very little perturbed at the initial instant, then ω x t 0 and ω y t 0 are very
small, that means that C 1 and C 2 are also very small (solutions of the system composed
of the equations (16.34) and (16.35). Since ω x and ω y are dependent on time by means
of the bounded functions sin p t and cos pt , it results that these components of the angular
velocity remain very small during the motion as compared to the component ω z = ω 0

supposed to be very large. It results that the vector ω does not deviate too much from the
initial unperturbed position; hence the gyroscope motion is stable. This property made the
gyroscope to be used in practice as a compass and as a stabilizer.

16.2.4.3. Gyroscopic effect

Let us consider a centered gyroscope (Figure T 16.4) having the axis Oz of the

movable trihedral as axis of rotation. Denoting by ω the (very large) angular velocity of
the gyroscope and by J the moment of inertia about the axis of rotation, then the angular

momentum K O of the gyroscope will be directed on Oz and will have the

magnitude K O = J ω . Acting on the gyroscope with a force F parallel to Oz, this produces

a moment M O relative to point O having direction of the axis Oy.

This moment produces a variation of angular momentum K O , variation given by
the theorem of the angular momentum:

d KO → → →
=M O ⇔ d KO = M O dt (16.38)
dt

278
Theoretical mechanics – theory and applications


F 
M0  y

K′0 dK 0
 
ω O≡C dϕ K0 z

x
Figure T 16.4
→ →
Vector d K O is collinear with M O . The new value of the angular momentum is:
→ → →
'
KO = K O+ d K O (16.39)

'
The line of action of the vector K O is the new axis of rotation of the gyroscope. It
is inclined with respect to the initial rotation axis Oz with a small angle d ϕ due to the
stability property of gyroscope.
Consequently, when acting the centered gyroscope with a force, then the axis of rotation
undergoes a small displacement in a plane perpendicular to the direction of the force.
This phenomenon is called gyroscopic effect.

16.3. Dynamics of the plane-parallel motion


Let us consider a rigid in a plane-parallel motion acted upon by external forces that
→ → 
are reduced about the center of mass of the rigid C to the torsor τ C  R, M C  .
 
Motion study (Figure T 15.5) is performed with respect to a fixed reference O 1 x 1 y 1 z 1
having the plane O 1 x 1 y 1 parallel to the fixed plane related to which the motion takes place
(see the " Kinematics of the plane-parallel motion") and that contains the center of mass
(C ∈ O 1 x 1 y 1 ) and a movable reference Oxyz, stiff-connected to the rigid, having the origin
at the center of mass ( O ≡ C ) .
At an arbitrary instant of time t, the coordinates of the center of mass related to the fixed
( )

∧ 



∧ 
reference are C x 1C , y1C ,0 and mOx, O 1 x 1  = mOy, O 1 y 1  = θ .

z1 O1 z1

z1 θ R
z 
 z
MC
MC  y
R
O1 y1C y1
O≡C O1 y1
y O1 x1
x1C θ ψ ϕ x
O1
O≡C O1y1
N
y1
x1
O1x1
θ x1
x
Figure T 16.5 Figure T 16.6

279
Theoretical mechanics – theory and applications

The rigid position is determined by three independent scalar parameters, namely:


x 1 C = x 1C ( t ) , y 1C = y 1C ( t ) , θ = θ (t)

In order to determine these unknown functions the theorems of the motion of the
center of mass and of the angular momentum about the center of mass are applied.

Theorem of the motion of the center of mass:


→ →
M a C= R (16.40)
is projected on the axes of the fixed reference system. Equations obtained are as follows:
•• ••
M x 1C = R x , M y 1C = R r , 0= R z (16.41)
The projections of the vector equation representing the theorem of the angular
momentum about the center of mass C on the axes of the movable reference have the same
form as with the rigid with fixed axis:
•• • •• • ••
− J C x z θ + J C y z θ 2= M Cx, − J C y z θ − J C x z θ 2= M Cy , JC z θ = MC z (16.42)
→ → → →
where M C = M Cx i + MC y j + MCz k .
The initial conditions for integration of equations (16.41) and (16.42) relate to the
position and velocity at the initial instant t 0 :
 x (t ) = x 0 , y (t ) = y 0 , θ (t ) = θ
 1C 0 1C 1C 0 1C 0 0
t = t 0 : • • • (16.43)
 x 1C (t 0 ) = v C0 x , y 1C (t 0 ) = v C0 y , θ (t 0 ) = ω 0

Particular case: In the case of a flat plate moving in its plane the motion study is
performed with respect to a fixed reference O 1 x 1 y 1 (chosen in the motion plane) and with
respect to the movable reference Cxy, stiff-connected to the rigid with the origin at the
center of mass C and the axes Cx and Cy as principal axes and central axes of inertia. The
equations of motion obtained by customizing the equations (16.41) and (16.42) are:
•• •• ••
M x 1C = R x , M y 1C = R y , J Cz θ = M Cz (16.44)
If the rigid is free, the parameters x 1C , y 1C , θ are independent, with no
relationship between them. The equations left available in the system (16.44) serve
determining the reactions.

16.4. Rigid dynamics in the general motion

Let us consider a rigid in general motion subject to action of an external force


→ → 
system reduced about the center of mass C to the torsor τ C  R , M C  . The motion study is
 
 
performed with respect to the fixed reference O 1 x 1 y 1 z 1 and the movable reference Oxyz,
originating in the center of mass ( O ≡ C ) with the axes Ox, Oy and Oz as principal axes
and central axes of inertia (Figure T 16.6).

280
Theoretical mechanics – theory and applications

The rigid position is fixed at an instant t, by means of the coordinates


x 1C , y 1C , z 1C of the center of mass with respect to the fixed reference and Euler’s
angles ψ , ϕ , θ . The six parameters of motion correspond to the six degrees of freedom of
the free rigid. The following equations result:
•• •• ••
M x 1C = R x , M y 1C = R r M z 1C = R z
J 1 ε x + ( J 3 − J 2 ) ω z ω y = M C x , J 2 ε y + ( J 1− J 3 ) ω x ω z = M Cy , (16.45)
J 3 ε z + (J 2− J 1 )ω yω x = M Cz ,

where J 1, J 2 , J 3 are the principal moments of inertia, R x , R y , R z the projections of the



resultant force R on the axes of the fixed reference and M C x, M C y , M C z the projections

of the resultant moment M C on the axes of the movable reference.
The initial conditions of the motion relate to the position and velocity at the initial
instant:
 x (t ) = x 0 , y (t ) = y 0 , z (t ) = z 0
1C 0 1C 1C 0 1C 1C 0 1C

 ψ (t 0 ) = ψ 0 , ϕ (t 0 ) = ϕ 0 , θ (t 0 ) = θ 0
t = t 0 : • • •
 x 1C (t 0 ) = v C x , y 1C (t 0 ) = v Cy , z 1C (t 0 ) = v C z
0 0 0 (16.46)

• • • • • •
ψ (t 0 ) = ψ 0 , ϕ (t 0 ) = ϕ 0 , θ (t 0 ) = θ 0

Solving the system (16.45) in the initial conditions (16.46) allows obtaining of the
unknown functions:
x 1C = x 1C ( t ) , y 1C = y 1C ( t ) , z 1C = z 1C ( t )
ψ = ψ (t) , ϕ = ϕ (t) , θ = θ (t) .

16.5. Solved problems

R 16.1) The homogeneous cylindrical body 1, of mass m = 50 kg, radius r = 15 cm and


height h = 2r, is stiff-connected to the shaft 2 of negligible mass, that can be rotated
around the horizontal axis AB (Fig. R 16.1 .1). The symmetry axis of the cylinder C z C is
forming the angle α = 15 0 with the rotation axis AB ≡ Cz. The shaft is acted upon by a
couple of forces located in a plane perpendicular to the shaft axis, of a constant
moment M = 15 N ⋅ m .
Knowing the distance AB = 2h = 4r = 60 cm, knowing that the system starts from
rest and neglecting the friction in the bearings A and B, determine:
a) The law of motion of the cylindrical body 1;
b) The reactions in the bearings A and B at the instant t 1= 2 s .

281
Theoretical mechanics – theory and applications

h = 2r h = 2r
y
yc


z M
B α
C
zc A

r
r x = xc
H

Figure R 16.1.1

Solution: Since the cylindrical body 1 has a motion of rotation the only degree of
freedom is the angle θ formed by the plane Cyz with itself between instants t = 0 and t =
→ →
arbitrary. We denote by R A and R B the reactions in the bearings A and B and apply the
theorem of the angular momentum about point A:

→ → → → → →
K = M + AC × G + AB× R B (1)
→ → →
Since the reactions R A and R B in the bearings A and B and the weight m g do
not give a moment in the direction z, equation (1) is rewritten as:
••
J z⋅ θ = M (2)
From (2), by successive integrations, we find that:
•• M ε t2
ε =θ = , ω = ∫ ε dt = ε t , θ = ∫ ω dt = (3)
Jz 2
The constant of integration are null due to the initial condition: t = 0, ω = θ = 0 .
The value of the moment of inertia will be determined later.
b) In order to determine the reactions R

A (R A x, R A y , R Az ) and →
( )
R B R B x , R B y ,0 the
theorem of the center of mass motion is applied:
→ → → →
m a C = R A+ R B + m g (4)
and the theorem of the angular momentum about the center of mass C:

→ → → → → →
K C = M + CA × R A + CB × R B (5)

→ →
As the center of mass C is on the fixed axis AB its acceleration is null ( a C = 0 ).
Projecting equation (4) on the axes of the movable reference system Cxyz the following
scalar equations are obtained:

0 = R Ax + R B x , 0 = R Ay+ R B y , 0 = R A z − mg (6)

282
Theoretical mechanics – theory and applications

The derivative with respect to time of the angular momentum of the cylindrical
body 1, with Cz as fixed axis, has the expression:


K C= − J ( xzε +J y zω
2 ) i + (− J

yzε −J x zω
2 ) j+ J

z

εk (7)
Simultaneously:
→ → →
i j k
→ → → →
CA × R A = 0 0 − h = h R Ay i − h RA x j
R Ax R Ay R Az
→ → →
i j k
→ → → →
CB × R B = 0 0 h = −h R B y i + h RB x j (8)
R Bx R By 0

Projecting equation (5) on the axes of the reference Cxyz and using (7) and (8) we
find the scalar equations:
− J x z ε + J y z ω 2 = h R A y − h RB y , − J y z ε − J x z ω 2 = −h R A x + h RB x , J z ε = M (9)
Obviously, the last equation (9) coincides with equation (1).
Solving the system composed of the equations (6) and the first two equations (9)
we obtain the projections of the reactions in the bearings A and B:
J y zε + J x zω 2 − J xzε + J y zω 2
RAx = −R B x = , R A y = −R B y = (10)
2h 2h
The only elements that still need to be determined are the centrifugal moments
J x z , J y z and the axial moment of inertia J z . In order to obtain them we determine first
the axial moments of inertia of the cylinder about the axes of the reference system
C x C y C z C stiff-connected to the body, which are principal axes of inertia.
According to the result found in solving the problem R 14.2 we have:
mr2
J zC = (11)
2
Due to the symmetry we also have:
J xC = J y C (12)
where
J xC = ∫∫∫ (y C2 + z C2 ) dm (13)
Denoting by ρ the density of the material from which the cylinder was made we
find that:
m m
dm = ρ dV = dV = dV (14)
V 2π r 3
In order to compute the triple integral in the relation (13) (on the entire domain
occupied by the cylindrical body) the system of cylindrical coordinates ( ρ , θ , z C ) shown
in Figure R 16.1.2 is used.

283
Theoretical mechanics – theory and applications

As
y C = ρ sin θ , z C = zC , dV = ρ dρ dθ d z C (15)
from (13) we obtain:
r 2π r
J xC =
m
3 ∫ ∫ ∫
( ρ 2 sin 2 θ + z C2 ) ρ dρ dθ dzC =
7
mr2 (16)
2π r 0 0 −r 12
zc
r
yc


dz c
r
ρ + dρ
M
θ
zc
yc
 M β
C xc

xc r

Figure R 16.1.2

The tensor of the moments of inertia of the cylinder about the reference system
C x C y C z C is:
J 0 0  7 0 0
 xC  mr2  
[J C ] =  0 J y C 0  = 0 7 0 (17)
12 
 0
 0 J z C   0 0 6 

The tensor of the moments of inertia [J] about the system Cxyz has the expression:
[ J ] = [α ] ⋅ [J C ] ⋅ [α ] T (18)
in which
1 0 0 
[ α ] =  0 cos α − sin α 

(19)
 0 sin α cos α 
is the matrix of transfer from the system C x C y C z C to the system Cxyz.
From (17), (18) and (19) it results:
 J − Jxy − J xz  7 0 0 
 x  mr 2  
[ J ] = − J y x Jy − J yz  =  0 6 + cos α
2
sin α cos α  (20)
12
 − J z x − Jzy J z   0 sin α cos α 6 + sin 2 α 
thus:

J xz = J zx = 0 , J yz= Jzy =−
mr 2
12
sin α cos α , J z =
mr 2
12
6 + sin 2 α ( ) (21)

284
Theoretical mechanics – theory and applications

Using the numerical data in the problem statement the following values are
obtained:
J x z = 0 , J y z = −0.023 kg ⋅ m 2 , J z = 0.568 kg ⋅ m 2
ε = 26.37 rad / s 2 , ω = 26.37 t rad / s , θ = 13.18 t 2 rad
R A x = − R B x = −2.02 N , R A y = − R B y = −106.6 N , R A z = 490 N .

R 16.2) A ball of weight G and radius r is thrown from a tower of height h at the horizontal
→ →
speed v 0 . At the instant of throwing the body is given an angular velocity ω 0 in a
direction perpendicular to the plane of motion (Figure R 16.2). Determine:
a) Laws of plane-parallel motion of the ball;
b) The time after which the ball touches the ground;
c) Distance from the tower base to the point of impact with the ground;
d) The speed at which the ball hits the ground;
e) The number of rotations performed by the ball in the air.
y1
O1 y1
y

ω0
θ

A=C V0

 C( x1c , y1c )
G 
h Vc x

G
x1
O1 B

Figure R 16.2

Solution: a) For the study of motion let us consider a fixed reference


system O 1 x 1 y 1 , having the axis O 1 y 1 superimposed on the axis of the tower and passing
through the center of mass C in the initial position of the ball, and a movable reference
system Cxy, stiff-connected to the rigid (see Figure R 16.2).
 ∧ 
Denoting by θ = masO 1 x 1, C x  and with x 1C , y 1C the coordinates of the center
 
of mass C about the fixed system, the system of differential equations of the motion is:
G •• G •• ••
x 1C = 0 , y 1C = −G , J C z θ = 0 (1)
g g
where Cz is the perpendicular at C to the plane Cxy.
By successive integrations with respect to the time t it is obtained:
• • •
x 1C = C 1 , y 1C = − g t + C 2 , θ =C3 (2)
and
gt2
x 1 C = C 1 t + C 4 , y 1C = − + C 2 t +C 5 ,θ =C 3 t +C 6 (3)
2

285
Theoretical mechanics – theory and applications

The initial conditions of the motion:


 x 1C = 0 , y 1C = h , θ = 0

t = 0 : • • • (4)
 x 1C = v 0 , y 1C = 0 , θ = ω 0

allow obtaining of the constants of integration C i , i = 1,6 :


C 1= v 0 , C 2 = 0 , C 3 = ω 0 , C 4 = 0 , C 5 = h , C 6 = 0 (5)
The parametric equations of plane-parallel motion of the ball are:
gt2
x 1C = v 0 t , y 1C = − +h , θ =ω0 t (6)
2
b) The ball touches the ground when y 1C = r , that is for:
2 ( h − r)
t 1= (7)
g
from the beginning of the motion.
c) Distance from the tower base (the vertical of the point of throw) to the point of impact
with the ground is:
2( h − r )
O 1 B = v 0 t 1= v 0 (8)
g
d) Velocity of the center of mass of the ball at an arbitrary instant has the components:
• •
v C x = x 1C = v 0 , v C y = y 1C = − g t (9)
At the instant of impact with the ground we may write that:

vC x 1 = v 0 , v C y 1 = − g t 1= − 2 g ( h − r ) , v C 1 = v 02 + 2 g ( h − r ) (10)

The velocity of a ball point located at its periphery (hence of the one that touches
the ground) is composed of the velocity of the ball center and velocity of rotation around
the ball center:

v r= ω 0 r (11)

Since v r is horizontal and directed in the opposite sense to the positive sense on
Ox, the velocity the ball touches the ground at has the components:

v x 1 = v 0− ω 0 r , v y 1 = − 2 g ( h − r) (12)

and the magnitude v 1= (v 0 − ω 0 r ) 2 + 2 g ( h − r ) .


e) The angle the ball rotated with during its motion in the air is equal to
2 ( h − r)
θ 1= ω 0 t 1= ω 0 (13)
g
hence the number of rotations performed by the ball is:

286
Theoretical mechanics – theory and applications

θ 1 ω0 h−r
N= = (14)
2π π g

R 16.3) A train locomotive describes a curve of radius R with the velocity v of its center of
mass. Knowing that the weight of the locomotive is G, the distance between the wheels is l
and the wheel radius is r, specify how much increase or decrease the reactions N A and
N B (Figure R 16.3).
z1 z1
l z B

 RB
V θ l /2
 y
z ω1 ⊗

  C ⊗   O y1
Mg
ω2 r F −F 
x1  ω l /2
A B
O1 x RA
R x1 A
 
NA NB

Figure R 16.3 Figure R 16.4

Solution: Before entering the curve, the two reactions are equal in magnitude:
G
N A= N B = (1)
2
In the curve, the wheels-axle assembly forms a gyroscope with horizontal axis
having an own rotation of angular velocity:
v
ω 1= (2)
r
and a precession motion of angular velocity:
v
ω 2= (3)
R
Denoting by J z the moment of inertia about the axis of inertia Cz and noting
 ∧  → → →
that θ = masC z , O 1 z 1  = 90 , we find that the gyroscopic moment M g = J z ω 1 × ω 2 has
 
the direction and sense of the motion of the train, increasing the pressure on the external
rail and decreasing the one on the internal rail by the same force:
M g J z ω 1ω 2 J z v 2
F= = = (4)
l l Rrl
Reactions N A and N B in the curve shall have the values:
G G J z v2 G G J z v2
N A= +F= + , N B= −F= − (5)
2 2 Rrl 2 2 Rrl

287
Theoretical mechanics – theory and applications

R 16.4) A body of revolution with the axis of symmetry Oz, rotates around the fixed axis
O z 1 that passes through the center of gravity of the body (Figure R 16.4). Knowing the

constant angular velocity ω of the body, angle θ between Oz and O z 1 , distance l
between the bearings A and B and the moments of inertia J x and J z of the body,
determine the reactions in the bearings.
Solution: Since point O belongs to the fixed axis O z 1 its velocity is zero, hence
the revolution body is treated as a rigid with fixed point. Choosing axis Ox in the plane
→ →
O z z 1 we obtain the following projections of the vectors ω and ε :
ω x = −ω sin θ , ω y = 0 , ω z = ω cos θ , ε x = ε y = ε z = 0 (1)
→ →
The only forces giving a moment about the point O are the reactions R A and R B .
Euler’s equations for this problem lead to:
M x = M z = 0 , M y = − ω 2 (J x − J z ) sin θ cos θ (2)
ω2
M O= M y= ( J x − J z ) sin 2θ (3)
2

The moment M O shall be directed along the axis Oy and shall be produced by the
components R A x and R B x (which form a couple of forces). Its magnitude is:
M O = R Ax⋅ 2 l (4)
hence:
ω 2(J x − J z ) sin 2θ
R Ax = −R B x = (5)
4l
The other components of the reactions in A and B are zero.

  ∧  
R 16.5) A homogenous bent bar ABC  m ABC  = 90 0  is of weight 3G and is hinged in
   
the point B (Figure R 16.5.1). The bar is attached at A by a spring of elastic constant k. At
equilibrium the side BC forms the angle α with the horizontal direction. Determine the
angle α so that the period of the small oscillations is minimum. What is the value of this
period?

B B
⋅  C1 ⋅ α
α Fel
k •
a
 C
A A
G • 2
2a
 θ
2G

C
C

Figure R 16.5.1. Figure R 16.5.2

288
Theoretical mechanics – theory and applications

Solution: Let us consider the bar in an arbitrary position defined by the angle θ related to

the position of equilibrium (Figure R 16.5.2). It is acted upon by the forces of weight G

and 2 G (at the centers of mass of the bar segments AB and BC, respectively) and the

elastic force F el (perpendicular to AB at A), of magnitude F el = F el static
+ k ⋅ a θ . F el
static

represents the value of the elastic force at the equilibrium of the bar. It is determined from
the equation of moments about the point B (see Figure R 16.5.1, for θ = 0 ):
a
static
F el ⋅ a + G ⋅ sin α − 2 G ⋅ a cos α = 0 (1)
2
It results:
 1 
static
F el = G  2 cos α − sin α  (2)
 2 
For the study of the small oscillations of the bar, the theorem of the angular
momentum is used projected on the axis ∆ , perpendicular to the bar plane at B:
••
J ∆⋅ θ = M ∆ (3)
wherein:
G a 2 2G ( 2a) 2 3 G a 2
J∆ = + = (4)
g 3 g 3 g
a
M∆ = 2G ⋅ a cos( α + θ ) − G ⋅ sin ( α + θ ) − F el ⋅ a (5)
2
But
sin ( α + θ ) = sin α cos θ + sin θ cos α ≅ sin α + θ cos α (6a)
cos( α + θ ) = cos α cos θ − sin θ sin α ≅ cos α − θ sin α (6b)

since for the small oscillations of the bar we may consider sin θ ≅ θ and cos θ ≅ 1 . With
the relations (6), the equation of motion of the bar becomes:
•• g  1 ak
θ+  2 sin α + cos α + ⋅θ = 0 (7)
3a  2 G 
The pulsation of the small oscillations has the expression
g  cos α a k 
p=  2 sin α + +  and the period of the small oscillations:
3a  2 G 
2 π not
T= = f (α ) (8)
p
As
df 2π d p πg  sin α 
=− 2 =−  2 cos α −  (9)
dα p dα 3a p 2 2 
The condition for obtaining of the extreme value of the period of the small
d f
oscillations = 0 allows obtaining the value sought for the angle α :

289
Theoretical mechanics – theory and applications

α min = arctg 4 (≅ 75,9 0 ) (10)


It can be shown that this value corresponds to a minimum of the function f ( α ) .
This is:

Tmin = f (α ) = 2π 6a G
min
(
g 2 a k + G 17 ) (11)

16.6. Problems proposed

16.6.1. Conventional tests



TC 16.1) A homogenous rod OA of length l and weight G can rotate around the horizontal
axis passing through the point O (Figure TC 16.1.1). Initially the bar is at rest in horizontal
direction. Leaving it free, it moves under its own weight.
Knowing that the moment of friction in joint O is M f , determine:
a) Angular acceleration ε depending on the position θ of the bar;

b) Reaction R O in joint O when the bar is in vertical position;
c) The angle β corresponding to the extreme position of the bar.
y1

B(µ)

O θ
β A C l
l ,G

mg


θ0 V0
O1 x0 A (µ) x1

Figure TC 16.1.1 Figure TC 16.2.1

TC 16.2) A bar AB, of mass m and length l, is moving with friction in a vertical plane so
that its ends are continuously resting on two walls, one vertical and one horizontal (Figure
TC 16.2.1). The coefficient of friction between bar and walls is µ . At the instant t = 0 the

end A was at the distance x 0 to the vertical guide and its velocity was v 0 . Determine the
differential equation of bar motion.

TC 16.3) A disc of weight G and radius r rolls without slipping on an inclined plane at the
angle α with the horizontal direction (Figure TC 16.3). The coefficient of rolling friction
is s. Find the laws of motion if at the initial instant the disc starts from the height h without
initial speed.

290
Theoretical mechanics – theory and applications

TG 16.1) Which of the following formulae expresses the theorem of the angular
momentum in the absolute motion of a material point?
 
dH     dKO  1
a) = F ; b) K 0 = r × H ; c) = M 0 ; d) E = m v 2 .
dt dt 2
y
G, r
XC
θ
C
T=?
r
h
C
A(s) xc = ?

G

x
α

Figure TC 16.3 Figure TG 16.2


TG 16.2) A disc of radius r and weight G is wrapped by an inextensible cord, of
negligible weight, of which upper extremity is fixed to a beam (Figure TG 16.2). If at the
initial instant the disc is against the beam, find the tension in the cord during the disc
motion.
G G G
a) T = G ; b) T = ; c) T = ; d) T = .
2 3 4

16.7. Indications and answers

TC 16.1) See Figure TC 16.1.2.


a) The theorem of the angular momentum is applied about the fixed point O:
l
J O ε = G ⋅ cos θ − M f (*)
2
G l 2
3  2M f 
 g .
As J O = , we find that ε =  cos θ −
g 3 2l Gl 
G→ → →
b) The theorem of center of mass motion a C = G + R O , is projected on Ox and Oy: We
g
get:
G
R Ox= a C x=
g
G
g
(
− a C τ sin θ − a C ν cos θ )
and
G
R O y= G + a C y= G +
g
G
a
g Cτ
(
cos θ − a C ν sin θ )
lε lω 2
and a C τ = and a C ν = , respectively.
2 2

291
Theoretical mechanics – theory and applications

The square of the angular velocity, ω 2 , is obtained by applying the work-kinetic


energy theorem:
1 l
J O ω 2 − 0 = G ⋅ sin θ − M f ⋅ θ
2 2
3 2M f 
and the angular acceleration from the relation (*). We have ω 2 = sin θ − θ  g .
l Gl 
π
For θ = we get:
2
G G G l 3M f
RO x = a C x = − aC τ = − ε=
g g g 2 2l
G G G l 2 3  π M f 
RO y = a C y = − aC ν = − ω =  − G  .
g g g 2 2 l 

c) The work-kinetic energy theorem is applied between the initial instant (the bar is in
horizontal position) and the instant of stop:
0 − 0 = G ⋅ sin β − M f (π − β )
l
2
The angle β shall be the solution of the equation:
2M f 2π M f
sin β + β = .
Gl Gl

R0 y1
y

TB
z  
• R0 NB
x B
 x
β  a cν θ
Mf C(x1c , y1c )
 
a cτ A
y G  
mg
NA
θ
O1  A
TA x1

Figure TC 16.1.2 Figure TC 16.2.2

TC 16.2) Release the bar from its constraints (friction supports in A and B) and introduce
→ → → →
the corresponding forces N A , T A , N B and T B (Figure TC 16.2.2). The system of
differential equations of the bar plane-parallel motion is:
•• ••
m x 1C = N B − T A , m y 1C = N A + T B − m g
•• l l l l
J ∆C⋅θ = − N A⋅ cos θ + T A⋅ sin θ + N B ⋅ sin θ + T B ⋅ cos θ
2 2 2 2

292
Theoretical mechanics – theory and applications

ml2
to which we add the relations: T A = µ N A , T B = µ N B , J ∆ C = .
12
The coordinates x 1C , y 1C , θ are related by the relations:
l l
x 1C = cos θ , y 1C = sin θ
2 2
We obtain thus a system of seven equations and seven unknowns:
x 1C , y 1C , θ , N A , T A , N B , T B . Eliminating the normal reactions and the friction forces we
obtain the following second order inhomogeneous non-linear differential equation with the
unknown θ ( t ) :

••
θ+
3µ •
θ 2=
6µ g
sin θ −
(
3 1− µ 2 g
cos θ
)
2− µ 2
2− µ 2 l 2− µ 2 l

The initial conditions of the motion are:


• x0
t = 0 : x 1C = v 0 , θ 0 = arccos .
l

TC 16.3) A fixed reference system Cxy is used, with the axis Cx parallel to the inclined
plane and with the origin at the initial position of the center of mass C. The system of the
differential equations of motion is:

0 = N − G cos α
 G ••
 x C = G sin α − T (1)
g
 ••
 J ∆ C ⋅ θ = T ⋅ r − M r

where M r = s N . The fact was used that y C = constant (the disc center motion is parallel
to the inclined plane). Between the disc center velocity and the angular velocity of the disc
plane-parallel motion there is the relation vC = r ω (the ICR is at the point of contact of
• • •• ••
the disc with the inclined plane). We find thus that x C = r θ and x C = r θ . Under the initial
conditions:

 x C = θ = 0
t = 0 : • • (2)
 x C = θ = 0

by integrating the system of differential equations (1), we obtain the following laws of
motion of the disc:

293
Theoretical mechanics – theory and applications

 g s 
 x C = sin α − cos α  t 2
 3 r 
y C= 0 (3)

g
θ = ( r sin α − s cos α ) t 2
 3r 2

TG 16.1) Theoretical result. Correct answer: c)

TG 16.2) The disc has a plane-parallel motion. The theorem of the center of mass motion
and the theorem of the angular momentum are applied about point C:
••
m xC = G − T , JC ⋅ ε = T ⋅ r
••
x G r2 G
But ε = and J C = , so that T = . Correct answer: c).
r g 2 3

294
Theoretical mechanics – theory and applications

BIBLIOGRAPHY

1. Brândeu L., Orgovici L., Chioreanu M., Teoremele generale ale dinamicii, Culegere de
probleme, Timişoara, 1991.
2. Ceauşu V., Enescu M., Ceauşu F., Culegere de probleme de mecanică, Institutul
Politehnic Bucureşti, 1984.
3. Dumitraşcu Ghe., Deleanu D., Mecanică teoretică, Editura ExPonto, Constanţa 1998.
4. Deleanu D., Dumitraşcu Ghe., Seminarii de mecanică, Editura Printech, Bucureşti,
2002.
5. Deleanu D., Dumitraşcu Ghe., Statica, Culegere de probleme, Editura Printech,
Bucureşti, 2000.
6. Deleanu D., Dumitraşcu Ghe., Cinematica, culegere de probleme, Editura Printech,
Bucureşti, 2000.
7. Deleanu D., Dumitraşcu Ghe., Dinamica, Culegere de probleme, Editura Printech,
Bucureşti, 1999.
8. Iacob C., Mecanică teoretică, Editura Didactică şi Pedagogică, Bucureşti, 1980.
9. Landau L., Lifşiţ F., Mecanica, Editura Tehnică, Bucureşti, 1966.
10. Niculescu M, Dinculescu N., Marcus S., Analiză matematică, Editura Didactică şi
Pedagogică, Bucureşti, 1966.
11. Olariu V., Sima P., Achiriloaie V., Mecanică teoretică, Editura Tehnică, Bucureşti,
1982.
12. Olariu V., Analiză matematică, Editura Didactică şi Pedagogică, Bucureşti, 1981.
13. Ripianu A., Popescu P., Bălan B., Mecanică teoretică, Editura Didactică şi Pedagogică,
Bucureşti, 1979.
14. Roşca I., Mecanică pentru ingineri, Editura MatrixRom, Bucureşti, 1998.
15. Roşculeţ M., Analiză matematică, Editura Didactică şi Pedagogică, Bucureşti, 1983.
16. Silaş M., Mecanică, Editura Didactică şi Pedagogică, Bucureşti, 1968.
17. Targ S., Theoretical Mechanics, Editura Mir, Moscova, 1975.
18. Tocaci E., Mecanică, Editura Didactică şi Pedagogică, Bucureşti, 1968.
19. Voinea R, Voiculescu D., Ceauşu V., Mecanică, Editura Didactică şi Pedagogică,
Bucureşti, 1983.

295
Tiparul executat în
Tipografia
UNIVERSITĂłII MARITIME
ConstanŃa

223

View publication stats

Potrebbero piacerti anche