You are on page 1of 126

MPH-002

CLASSICAL
Indira Gandhi National
Open University
MECHANICS-I
School of Sciences

Block

1
THE LAGRANGIAN FORMULATION OF MECHANICS
UNIT 1
Elementary Principles of Mechanics: A Revision 9
UNIT 2
Constrained Motion and the D'Alembert’s Principle 41
UNIT 3
Lagrange’s Equations and its Applications 69
UNIT 4
Hamilton’s Principle 99
Programme Design Committee
Prof. V.B. Bhatia, Retd. Prof. Enakshi Sharma Prof. G. Pushpa Chakrapani Prof. Suresh Garg, Retd.
University of Delhi, Delhi University of Delhi, South BRAOU School of Sciences,
Prof. Abhai Mansingh, Retd. Campus, Delhi Prof. Y.K. Vijay IGNOU, New Delhi
University of Delhi, Delhi Prof. H.S. Mani, Retd. University of Rajasthan, Prof. Vijayshri
Prof. Feroz Ahmed, Retd. IIT Kanpur Rajasthan School of Sciences,
University of Delhi, Delhi Prof. S. Annapoorni Prof. J. Nag, Retd. IGNOU, New Delhi
Prof Yashwant Singh, Retd. University of Delhi, Delhi Jadavpur University Prof. S.R. Jha
Banaras Hindu University, Prof. D. Choudhury Prof. Zulfequar, School of Sciences,
Varanasi University of Delhi, Delhi Jamia Milia Islamia, New Delhi IGNOU, New Delhi
Prof. Deepak Kumar Prof. T.R. Seshadri Dr. Om Pal Singh Prof. Shubha Gokhale
J.N.U., New Delhi University of Delhi, Delhi IGCAR, Kalpakkam, School of Sciences,
Tamil Nadu IGNOU, New Delhi
Prof. Vipin Srivastava Prof. S. Ghosh
Central University of J.N.U., New Delhi Prof. Prabhat Munshi Prof. Sanjay Gupta
Hyderabad, Hyderabad IIT Kanpur School of Sciences,
Prof. Neeraj Khare
IGNOU, New Delhi
Prof. G.S. Singh IIT Delhi, Delhi Prof. R.M. Mehra, Retd.
IIT Roorkee, Roorkee Dept. of Electronics, South Dr. Subhalakshmi Lamba
Prof. V.K. Tripathi
Campus, University of Delhi, School of Sciences,
Prof. A.K. Rastogi. IIT Delhi, Delhi
Delhi IGNOU, New Delhi
J.N.U., New Delhi Prof. Pankaj Sharan, Retd.
Prof. A. K. Ghatak Jamia Milia Islamia, Prof. S. K. Kulkarni Dr. M.B. Newmai
IIT Delhi, Delhi New Delhi Pune University/ School of Sciences,
IISER Pune, Pune IGNOU, New Delhi
Prof. Rupamajari Ghosh Prof. Kirti Ranjan
J.N.U., New Delhi University of Delhi, Delhi

Course Design Committee


Prof. V.B. Bhatia, Retd. Prof. G.S. Singh Prof. Neeraj Khare Prof. Vijayshri
University of Delhi, Delhi IIT Roorkee, Roorkee IIT Delhi, Delhi SOS, IGNOU, New Delhi
Prof. Abhai Mansingh, Retd. Prof. A.K. Rastogi Prof. Pankaj Sharan, Retd. Prof. S.R. Jha
University of Delhi, Delhi J.N.U., New Delhi Jamia Milia Islamia, New Delhi SOS, IGNOU, New Delhi
Prof. Feroz Ahmed, Retd. Prof. S. Annapoorni Prof. Kirti Ranjan Prof. Shubha Gokhale
University of Delhi, Delhi University of Delhi, Delhi University of Delhi, Delhi SOS, IGNOU, New Delhi
Prof. Yashwant Singh, Retd. Prof. D. Choudhury Prof. R.M. Mehra, Retd. Prof. Sanjay Gupta
BHU, Varanasi University of Delhi, Delhi Dept. of Electronics, South SOS, IGNOU, New Delhi
Prof. Deepak Kumar Prof. T.R. Seshadri Campus, Delhi University, Dr. Subhalakshmi Lamba
J.N.U., New Delhi University of Delhi, Delhi Delhi SOS, IGNOU, New Delhi
Prof. Vipin Srivastava Prof. S. Ghosh Prof. Suresh Garg, Retd. Dr. M.B. Newmai
Central University of J.N.U., New Delhi SOS, IGNOU, New Delhi SOS, IGNOU, New Delhi
Hyderabad, Hyderabad

Block Preparation Team


Dr.Subhalakshmi Lamba (Units 1-4)
School of Sciences, IGNOU, New Delhi
Course Coordinators: Dr. M. B. Newmai, Dr. Subhalakshmi Lamba
Block Production Team
Sh. Rajiv Girdhar
AR (P), IGNOU
Acknowledgement: Shri Gopal Krishan Arora, EDP, SOS for CRC preparation.
June, 2023
© Indira Gandhi National Open University, 2023
ISBN:
Disclaimer: Any materials adapted from web-based resources in this module are being used for educational purposes only and not for
commercial purposes.
All rights reserved. No part of this work may be reproduced in any form, by mimeograph or any other means, without permission in
writing from the Copyright holder.
Further information on the Indira Gandhi National Open University courses may be obtained from the University’s office at Maidan
Garhi, New Delhi-110 068 or the official website of IGNOU at www.ignou.ac.in.
Printed and published on behalf of Indira Gandhi National Open University, New Delhi by Prof. Meenal Mishra, Director, SOS, IGNOU.
Printed at
CONTENTS
Block and Unit Titles 1
Credit page 2
Contents 3
CLASSICAL MECHANICS-I: COURSE INTRODUCTION 5
BLOCK 1: THE LAGRANGIAN FORMULATION OF MECHANICS 7
Unit 1 Elementary Principles of Mechanics: A Revision 9
1.1 Introduction 9
1.2 Motion of a Single Particle 10
1.2.1 Dynamical Variables and Equation of Motion 12
1.2.2 Conservation Laws 14
1.3 Motion of a Many-Particle System 17
1.3.1 Dynamical Variables and Equation of Motion 17
1.3.2 Conservation Laws for a Many-Particle System ` 21
1.4 Summary 27
1.5 Terminal Questions 29
1.6 Solutions and Answers 30
Unit 2 Constrained Motion and the D'Alembert’s Principle 41
2.1 Introduction 41
2.2 Constraints 42
2.2.1 Examples of Constrained Motion 43
2.2.2 Classification of Constraints 45
2.3 D'Alembert’s Principle 49
2.3.1 Virtual Work 50
2.3.2 D'Alembert’s Principle 53
2.4 Generalized Coordinates 54
2.5 Summary 61
2.6 Terminal Questions 63
2.7 Solutions and Answers 65
Unit 3 Lagrange’s Equations and its Applications 69
3.1 Introduction 69
3.2 Lagrange’s Equations from D'Alembert’s Principle 71
3.2.1 Lagrange’s Equations of the Second Kind 71
3.2.2 Lagrangian 77
3.2.3 Kinetic Energy in Generalized Coordinates 81
3.3 Lagrange’s Equations for a Velocity-dependent Potential 82
3.4 Euler-Lagrange Equations of Motion for Simple Systems 85
3.4.1 Unconstrained Motion of a Single Particle 85
3.5 Summary 87
3.6 Terminal Questions 89
3.7 Solutions and Answers 90
Appendix 3A Calculus of Variations 97

3
Unit 4 Hamilton’s Principle 99
4.1 Introduction 99
4.2 Hamilton’s Principle and the Euler-Lagrange Equations 100
4.2.1 Configuration Space 100
4.2.2 Hamilton’s Principle 102
4.2.3 Derivation of the Euler –Lagrange Equations of Motion
from Hamilton’s Principle 103
4.2.4 Generalized Momentum 106
4.2.5 Energy Function 108
4.3 Conservation Theorems 109
4.3.1 Cyclic Coordinates and Conservation of Generalized Momentum 109
4.3.2 Conservation of the Energy Function 110
4.3.3 Integrals of Motion 110
4.4 Symmetry-Homogeneity and Isotropy 111
4.5 Hamilton’s Principle for Nonholonomic Systems 116
4.6 Summary 119
4.7 Terminal Questions 121
4.8 Solutions and Answers 121

Further Readings 126


Table of Physical Constants 127
List of Blocks and Units: MPH-002 128
Syllabus: Classical Mechanics -I (MPH-002) 129

4
CLASSICAL MECHANICS I : COURSE INTRODUCTION
Mechanics, as you all know, is the study of the motion of material bodies: either a particle
or a system of particles. In modern times it has become customary to refer to “Mechanics”
as “Classical Mechanics” to distinguish it from the other important theory that developed in
the early years of the twentieth century: namely “Quantum Mechanics”. The physical
principles, on the basis of which the motion of objects under the influence of different kind
of forces are analysed, were laid down by Sir Isaac Newton, in a work entitled
Philosophiae Naturalis Principia Mathematica (Mathematical Principles of Natural
Philosophy) in 1687. The study of the motion of a particle (or a collection of particles)
using Newton’s
 Laws starts with the equation of motion for each particle:
 dv  
F m  ma  mr . The basic premise in being able to obtain the correct equation of
dt
motion and hence correct solution, is a knowledge of all the forces acting on the object.
But there are situations in which one knows only the effect of the force and not the force
itself. For example there are the forces that compel an object to move along a particular
path or surface . Similarly, one does not know the forces that keep the individual particles
in a rigid body at a fixed distance from each other. To use Newton’s laws we would first
have to calculate these forces and you have done that yourself while solving problems
though you may not have realized it. Newton’s laws are also not quite that straightforward
to implement when the system is not best described in the Cartesian coordinate system ( in
cartesian coordinates the equations of motion are simply Fx  mx; Fy  my; Fz  mz , not
always so in other coordinates).
Several years after Newton, Leonhard Euler (1707-1783), Jean le Rond d'Alembert
(1717-1783), Joseph-Louis Lagrange (1736-1813) and Sir William Rowan Hamilton (1805-
1865) successively reformulated mechanics using powerful analytical techniques. Not only
did these techniques make it possible to solve many problems in mechanics that were
somewhat intractable, they also have applications beyond mechanics. The Lagrangian is
considered the fundamental object which describes a quantum field theory and the
Hamiltonian determines the time evolution of the wave function in quantum mechanics. In
fact many of the concepts used in the formulation of quantum mechanics, already exist in
the Lagrangian and Hamiltonian formulation of mechanics.
In this course on Classical Mechanics-I we study the Lagrangian formulation of mechanics
and its applications. The Lagrangian formulation of mechanics works with scalar quantities
like kinetic energy and potential and does away with need to know all the forces in the
system.
In Block 1 of the course “The Lagrangian Formulation Of Mechanics” we derive the
Lagragian formalism of mechanics and apply it to simple problems. In Unit 1 we revise
revise the principles of Newtonian mechanics for the motion of a single particle and for
a many particle system by defining the dynamical variables , writing the equations of
motion and deriving the laws of conservation. In Unit 2 we discuss how the existences of
constraints in the dynamical system restricts the motion and introduce the notion of generalized
coordinates. You study the concept of virtual work which gives us the condition for the static
equilibrium of a dynamical system and the D’alembert’s Principle. In Unit 3 we define the
Lagrangian for a dynamical system, which is a scalar fuction and derive Lagrange’s equations ( or
the Euler-Lagrange equations) of motion and study some applications. It is also possible to derive
the Euler-Lagrange equations using a variational approach and we do this in Unit 4 . In this unit
you learn about the action of a dynamical system and state the Hamilton’s principle, also
5
called the Principle of Least Action. You will also study the conservation theorems and the
principles of symmetry.
In Block 2 of this course you will study the application of the Lagrangian formulation to two
very import problems in mechanics which have numerous applications in physics and
astronomy, namely the central force problem and small oscillations. In Units 5 and 6, you
will study two-body systems that deal with the motion of two particles under the influence of
their mutual gravitational attraction or some other central force. This problem has been
studied for centuries, and its solutions have deep implications for our understanding of
celestial mechanics and other physical phenomena. In Units 7 and 8, you will study small
oscillations, also known as harmonic oscillations, which refer to oscillatory motion around an
equilibrium position. In this type of motion, the restoring force is proportional to the
displacement from the equilibrium position, and the motion is periodic. Small oscillations are a
common phenomenon in many physical systems, including simple pendulums, mass-spring
systems, and the two-body problem.

In the end, we have a word of advice about how to study the course. You should not read this
course passively like a story but write down all concepts and work out all mathematical steps
given in the course on your own. Attempt all self-assessment questions (SAQs) and Terminal
Questions (TQs) given in the units. Do not skip any of those as they are designed to assess
your understanding of the subject.

We hope you enjoy studying the course. Our best wishes are with you!

6
BLOCK 1 : THE LAGRANGIAN FORMULATION OF
MECHANICS
In this block, we will introduce the basic concepts of the lagrangian formulation of mechanics
and its application to simple dynamical systems
In Unit 1 “Elementary Principles of Mechanics: A Revision” we present in brief the
principles of Newtonian mechanics as applied to single particle motion and motion in a
many particle system. We state the equations of motion and derive the laws of
conservation of linear momentum, angular momentum and mechanical energy. Most of
the matter in this Unit is familiar to you from your undergraduate courses and we have
provided many SAQs and TQS. Working through these will help you revise the concepts
and the methods of problem solving.
From Unit 2 onwards we extend our studies to include systems whose solutions are not
straightforward using Newton’s equation of motion . In Unit 2 “Constrained Motion and
the D'Alembert’s Principle”we describe motion in the presence of constraints, in which
the moving particle is restricted to move along a certain curve or surface. Physically, this
kind of constrained motion is because of the presence of “constraint forces” which
typically do not follow from a force law. The presence of constraints changes the number
of degrees of freedom of the dynamical system. You will learn about virtual work , the
“Principle of Virtual Work” which describes the static equilibrium of a constrained system
and the D’Alembert’s principle which describes the dynamics of a constrained system.
You also learn about generalized coordinates for a dynamical system.
In Unit 3 “Lagrange’s Equations and its Applications” we derive Lagrange’s equations of
the second kind, using D'Alembert’s Principle and the concept of generalized coordinates.
We define generalized velocity , momentum and force for the dynamical system. We
introduce an important scalar quantity called the “Lagrangian” for a dynamical system. Unlike
the Newtonian formulation in which we work with the force acting on a system, which is a
vector quantity. In the Lagrangian formulation, the equations of motion of the system are
derived from the Lagrangian. We derive the Euler-Lagrange equations, which are the
equations of motion for all the independent generalized coordinates and provide us with a
complete solution to any dynamical problem. You will work out the equations of motion for
several systems and you will see how the E-L equations generate the same equations of
motion that one derives from Newton’s second law of motion. The Lagrangian formulation
works with purely generalized coordinates, and generates directly the equations of motion of
the generalized coordinates, saving us the trouble of identifying forces, particularly the
constraint forces, which are vector quantities and transforming them from rectangular
Cartesian to any other coordinate system suitable to studying the dynamics of the system.
It is also possible to formulate dynamical problems in a whole new way, independent of
Newton’s law, using the “Variational Principle” which is an “integral” principle based on
the calculus of variations. In Unit 4 “Hamilton’s Principle”we introduce the idea of an n-
dimensional configuration space for a dynamical system with n degrees of freedom, in
which each point is labelled by n generalized coordinates and the evolution of the system
with time is represented by a curve in configuration space. We then introduce the action of
a dynamical system and state the Hamilton’s principle, or the Principle of Least Action, as it
is often called and derive the Euler Lagrange equations of motion for the system from the

7
Hamilton’s Principle. You will also study the conservation theorems which follow from the
E-L equations of motion

This block is mathematically intensive and it is expected that you are familiar with vector
algebra and differential and integral calculus. You should work through the steps of each
derivation and work out the examples, SAQs and Terminal Questions given in the units, to
understand the concepts better.

We wish you success!

8
Unit 1 Elementary Principles of Mechanics: A Revision

UNIT 1
ELEMENTARY
PRINCIPLES OF MECHANICS:
A REVISION
Structure

1.1 Introduction 1.3 Motion of a Many-Particle System


Expected Learning Outcomes Dynamical Variables and
1.2 Motion of a Single Particle Equation of Motion
Dynamical Variables and Conservation Laws for a Many-Particle
Equation of Motion System
Conservation Laws 1.4 Summary
1.5 Terminal Questions
1.6 Solutions and Answers

1.1 INTRODUCTION
In your undergraduate physics courses you have all studied the kinematics
and dynamics of different types of motion- motion along a straight line, motion
in a plane and rotational motion, analysed using Newton’s laws of motion.
Armed with a knowledge of the basic force laws and boundary conditions, you
can, in principle solve a second order differential equation (Newton’s second
law of motion) and determine the evolution of the motion of a particle or even
a collection of particles. It is also true that this method becomes progressively
more complicated, for example, when we are studying extended objects
instead of point particles.
Right up until the beginning of the twentieth century, it was believed that the
Newtonian description, is, in fact the most complete description of the motion
of all objects in the Universe. Today, however, we know that Newtonian
mechanics breaks down when the speed of the moving object is comparable
to the speed of light c. Such objects are studied using Einstien’s theory of
special relativity. It also breaks down when the objects do not have a
sufficiently large mass and/or kinetic energy. For the description of such
objects, for example, sub atomic particles such as electrons and protons, we
use quantum mechanics. Therefore modern science tells us that Newtonian
mechanics is an approximation and can be applied only to study the motion of
large objects moving with speeds that are small compared to the speed of
light.
9
Block 1 The Lagrangian Formulation of Mechanics
The fundamental principles for the study of motion, were completely specified
by Galileo and Newton in the 16th and 17th century. However, much later, in
the 19th century, the techniques of the study of motion were reformulated by
mathematical giants Euler, Lagrange, Hamilton and Jacobi, to name a few.
This analytical formulation of mechanics is what you will study in the courses
Classical Mechanics I and Classical Mechanics II. However, it is important to
understand, that while these sophisticated techniques help us work on
complex problems in mechanics and have applications beyond the study of
motion of objects, they do not add any new physical principles to the study of
motion.
In this Unit we briefly revise the principles of Newtonian mechanics that you
are already familiar with from your undergraduate courses. We begin with an
introduction to single particle motion in Sec. 1.2. We define the dynamical
variables and introduce the equation of motion and also the laws of
conservation. In Sec. 1.3 we describe in brief the motion of many particle
systems. Most of what you study in this Unit is familiar to you from your
undergraduate mechanics courses. We will not be giving you any worked out
examples in this Unit. Rather there is a collection of SAQs and TQS which will
test your understanding of these concepts and the methods of problem solving
that you are already familiar with.
In the next Unit we will start the formal mathematical formulation of mechanics
which is a prelude to the Lagrangian formulation.
Expected Learning Outcomes
After studying this unit, you should be able to:
 write down and solve the equations of motion for single particle motion;
 derive and apply the laws of conservation for single particle motion;
 derive the equations of motion for a many -particle system;
 obtain the centre of mass of a many particle system;
 derive expressions for the dynamical variables;
 derive the conservation laws for many particle systems; and
 solve problems based on the conservation laws for many-particle
systems.

1.2 MOTION OF A SINGLE PARTICLE


Let us start with a brief description the simplest type of motion that we are
familiar with : the motion of a single
 particle in three-dimensional space
under the influence of a force F . Before we do that, let us revise a few
concepts, which are important in the study of motion:
i) Newton’s Laws Apply to Particles

A particle is a concept that is central to mechanics. Essentially, a particle


is an object of insignificant size. But since we use the term to describe
objects ranging from tennis balls to planets, we should be a little more
precise. A particle is a body whose physical extent mat be neglected when
we describe its motion. Newton’s laws , as they were formulated, apply
only only to particles. So to study the motion of an extended object using
10
Unit 1 Elementary Principles of Mechanics: A Revision
Newtonian mechanics, we treat it as a particle. we assume that its entire
mass is concentrated at a single point, its centre of gravity. Clearly this is
an approximation and its validilty would depend on the problem at hand .
So while we can describe the motion of a planet around the sun by
assuming the planet to be a particle or a point object, in studying the
rotation of of the planet about its axis, it cannot be treated as a particle.
To apply Newton’s laws to an extended object, we treat its as a collection
of particles.

ii) Frame of Reference

All mechanical phenomena are described in a frame of reference. You are


also familiar with the notion that Newton’s laws of motion are valid only in a
certain class of frames of reference, which are called inertial frames of
reference . Newton’s first law of motion gives us a definition of an inertial
frame of reference. It is a frame of reference in which, in the absence of an
external force, a particle moves in a straight line with a constant speed. In
an inertial frame when a particle is at rest any one instant of time, it
remains at rest at all instants of time.

Since in an inertial frame, while moving in a straight line with constant


speed, the particle covers equal distances in equal intervals of time, both
space and time are homogeneous. Further since the straight path could be
along any direction, space is also isotropic. So we can say that in an
inertial frame of reference space is homogeneous and isotropic and
time is homogenous.

iii) Dynamical Variables

In general dynamical variables are defined as a set of variables that


completely specify the configuration of a mechanical system, that is the
position of all the particles in the system. These variables change with
time and under the action of forces. It is assumed that each of these
dynamical variables may change independent of the others( unless the
motion of the system is constrained in some way, as you will study in the
next Unit). So to specify the motion of the system we have to determine
how these dynamical variables change with time. This is done using
Newton’s second law.

Typically, the position of a particle in space is described by a position


vector relative to an origin. In the Cartesian coordinate system, this is
equivalent to specifying three position coordinates, x, y and z. For an n-
particle system, you would have n position vectors, one for each particle or
equivalently, 3n position coordinates.
iv) Degrees of Freedom
The number of degrees of freedom of a mechanical system, is the number
of independent quantities that are required to uniquely specify the position
of all the particles in the system. A single particle, if its motion not
restricted in any way, has three degrees of freedom. For an n-particle
system, the number of degrees of freedom would similarly be 3n.
11
We may choose to specify the degrees of freedom in different ways:
Block 1 The Lagrangian Formulation of Mechanics
 you may choose an origin relative to which the position of the system is
to be specified;
 you may choose different coordinate systems: Cartesian, cylindrical or
spherical polar.
We now describe the dynamical variables and the laws that govern the
evolution of these variables for single – particle motion.
1.2.1 Dynamical Variables and Equation of Motion
The position of a
particle in an n- If the position vector of the particle at an instant of time t ,with respect to a
dimensional Euclidean 
given origin is denoted by r (t ) , we can write the velocity of the particle
space is expressed by   
  dr  dv d 2 r 
an ordered set of n as: v  r  and its acceleration as a    r . If the mass of the
independent real dt dt dt 2
numbers, each of particle is m, its linear momentum is written as
which defines a  
position coordinate. At p  mv (1.1)
any instant of time, at
As you all know, in an inertial frame of reference, the equation of motion for
any given point in
space we can draw 3 the particle can be written using Newton’s second law as:
mutually perpendicular  dp d mv 
axes, so the dimension F  (1.2)
dt dt
of space should be 3.
Note that Newton’s law emphasizes that the measure of the effect of a force is
the rate of change of linear momentum and not the acceleration. This law
introduces one of the most fundamental physical quantities – the linear
momentum. The linear momemtum is fundamental in the formulation of both
quantum mechanics and special relativity. It is only when the mass of the
system is a constant, that this equation reduces to the more familiar form:
 
dv  
F m  ma  mr (1.3a)
dt

This is, as you can see, a second order differential equation for r (t ) . In the

Cartesian coordinate system, r (t )  x (t )iˆ  y (t ) ˆj  z(t )kˆ reduces to the
following set of three differential equations :
mx  Fx ; my  Fy ; mz  Fz (1.3b)

Where Fx , Fy and Fz are the components of the force along the x, y and z
axes.You can solve this differential equation to get the path of the particle,
that is, its position vector at each instant of time. To solve the dfferential
equation, you need only to know the initial conditions of the motion of the
 
particle, i.e. the position and velocity ( r and r ) of the particle at an initial
time, say t  t 0 .

In general, the force F could be a function of position, velocity and time. Let
us see a few examples. If the particle is moving in the gravitational field of a
fixed, heavy object of mass M at rest, located at the origin, then the force
acting on the particle at any instant of time can be written as:
  GMm 
F (r )   r (1.4)
r3
12
Unit 1 Elementary Principles of Mechanics: A Revision

Where r is the position of the particle relative to the origin, in this case the

force is only a function of r . There is an implicit dependence on t through

r (t ) but no explicit dependence on t. On the other hand, if the heavy object
were itself moving, under the influence of some other force, then the force
would be:
 
F (r , t )  
  3

GMm  

r  R (t ) (1.5)
r  R (t )
 
Where R  R (t ) is the instantaneous position of the heavy object and the
force now depends explicitly on time. You can imagine that you are studying
the motion of a satellite around the earth and the earth is moving due to the
gravitational field of the sun.
A familiar example of a velocity dependent force is the Lorentz force. The
force on a charge q moving in a region with electric and magnetic fields is:
     
F (r , r, t )  q E (r , t )  q r  B(r , t ) (1.6)
   
So in general, it would be more accurate to write F  F (r , r , t ) .

We can also define the angular momentum of the particle. If the position of

 respect to a point O is given by r , then the angular
the particle with
momentum L of the particle about O is:
  
L r p (1.7)
 
The torque  acting on the particle, also called the moment of the force F ,
about the point O is :
  
  r F (1.8 )

We can derive an equation for  , which is analogous to the familiar form of
Newton’s second law of motion in Eq. (1.2). Taking the derivative on both
sides of Eq. (1.7) wrt to t we can write:

dL d   
r  p   d r  mv 

 (1.9)
dt dt dt
We can simplify the right hand side of this equation as:
d   
r  mv   dr  mv   r  d mv 
 
(1.10)
dt dt dt

dr 
The first term on the RHS of this equation is zero because  v is parallel to
 dt
 dr 
mv and therefore the cross product  mv  is zero. So we get:
dt
d    d   
r  mv   r  mv   r  F (1.11)
dt dt
Now Eq. (1.9) reduces to:

dL d     
 r  mv   r  F   (1.12)
dt dt
Or more familiarly written as

 dL
 (1.13)
dt 13
Block 1 The Lagrangian Formulation of Mechanics

SAQ 1
a) A block of mass M slides without friction down a fixed inclined plane which
makes an angle  with the horizontal. Calculate the acceleration of the
block.
b) A particle of mass m is falling freely under the action of gravity in a
medium. If the medium exerts a retarding force on the particle, which is
proportional to its velocity, determine the displacement and velocity of the
particle as a function of time, assuming that the particle starts from rest.
c) The position vector for a particle of mass 3.0 kg at any instant of time is

r  4.0 m iˆ  2.0 m ˆj . If its velocity at this instant of time is
 
v  5.0 m s 1iˆ  3.0 m s 1 ˆj and a force of F  8.0 N ˆj is acting on the
particle, calculate the angular momentum of the particle and the torque
about the origin.

From these dynamical equations we can derive important conservation laws,


as we shall see in the next section.
1.2.2 Conservation Laws
Let us first discuss two important conservation laws that arise from Eqs. (1.2)
and (1.13).
I) Conservation of Linear Momentum

From Eq. (1.2), for F  0 , we get:

dp 
 0  p  constant (1.14)
dt
In other words, in the absence of a force (at all times), the linear
momentum of the particle remains constant during the motion of the
particle, so the particle is in uniform motion. This is exemplified in
Newton’s first law where he introduces the concept of uniform motion as
motion in the absence of force. We say that the linear momentum is, in
this case, a conserved quantity.
Note also that Eq.(1.14) is a vector equation. So you can see that if for
example any one Cartesian component of the force is zero, then the
component of the linear momentum in that direction will be conserved. So
if Fx  0, p x  constant , and so on

II) Conservation of Angular Momentum



From Eq. (1.13), for   0 , we get:

dL 
 0  L  constant (1.15)
dt
In other words, in the absence of a torque(at all times), the angular
momentum of the particle is a constant of motion.
 Remember that
    
  r  F so   0 not only for F  0 but also for F  0 if the cross product
   
r  F  0 , or r is parallel to F , as is the case for a central force. Once
14 again remember that this is a vector equation. So if one component of the
Unit 1 Elementary Principles of Mechanics: A Revision
torque is zero, the corresponding component of the angular momentum is
a constant.
III) Conservation of Energy
Let us consider the motion
 of a particle along a path under the influence of
a net external force F . The work done in moving the particle from the
point A to the point B is:
B 
W AB   F .ds (1.16)
A

Remember that Eq.(1.16) is a line integral, ds is an infinitesimal
displacement along the path of motion and the work done is a scalar
quantity.If the mass of the particle is a constant then we can write
 dv  
F m and ds  vdt and Eq. (1.16) reduces to:
dt
B dv  1B d   1 B d (v 2 )
W AB   m .vdt   m (v .v )dt  m  dt (1.17)
A dt 2 A dt 2 A dt
 
If the velocity of the particle is v A at the point A and v B at the point B on
the path, we can write

W AB 
1
2

m v B2  v A2  (1.18)

For a particle moving with a velocity v , the kinetic energy of the particle
is:
1  1
T  mv .v  mv 2 (1.19)
2 2
With this definition, Eq. (1.18) is just:
W AB  TB  TA (1.20)

where T A and TB are respectively the kinetic energies of the particle at


the points A and B. Eq. (1. 20) is therefore is a statement of what is also
called the Work-Energy Theorem, that is:
The work done by the net external force acting on a body is equal to
the change in its kinetic energy.
Let us now suppose that the force acting on the body is a conservative
force. Recollect what have learnt about a conservative force in your
undergraduate courses. We know that

i) the work done by a conservative force, say F , between any two points
A and B is the same for any physically possible path between these
two points.
This is equivalent to the following statement:

ii) the work done by a conservative force, say F , along a closed path is
zero:
 
 F .ds  0 (1.21)

15
Block 1 The Lagrangian Formulation of Mechanics
The necessary and sufficient condition for the force to be conservative
is that the force can be expressed as the gradient of a scalar function

of position, customarily denoted by V (r ) . Let us see how.
Using Stoke’s Theorem we can write:
 

  

 F .ds     F .dS (1.22)
S

where S is the surface enclosing the path. Since dS is arbitrary,
Eq. (1.22) implies that:
     
  F  0 or F  V (r ) (1.23)
This scalar function is called the potential energy or simply the
potential. You are familiar with several such conservative forces, for
example the gravitational force and the electrostatic force. With this
definition, for a conservative force Eq. (1.16) reduces to
B
  

W AB   V (r ).ds  VA  VB (1.24)
A
where VA and VB are respectively the potential at the points A and B.
Comparing Eqs. (1.20) and (1.24), we can therefore write:
TB  T A  VA  VB  VA  T A  TB  VB (1.25)
So for a conservative force, the total mechanical energy E, which is
the sum of kinetic (T) and potential energy (V) of the particle, is a
constant of motion.
We can summarize the laws of conservation as follows:

1. If the total force acting on a particle is zero, the linear momentum is


conserved

2. If the total torque acting on a particle is zero, then the angular momentum
is conserved

3. If the forces acting on a particle are conservative, then the total mechanical
energy of the system is conserved

As you will see while studying this course, a very useful tool in understanding
mechanical systems is determining, apriori, the conserved quantities.
Conserved quantities are those that do not change with time. A conserved
quantity in mechanics is typically a function of position and momentum, say
 
Q(r , p ) which does not change when we evaluate it along the motion of the
particle. So we can write:
 
dQ( r , p )
0 (1.26)
dt

Later on in this course you will also study the connection between these
conservation laws and the fundamental symmetries in physics. In the words of
the famous theoretical physicist Robert L. Mills:

“For every conservation law, there is a symmetry.”


16
Unit 1 Elementary Principles of Mechanics: A Revision

SAQ 2
a) Determine which of the following is a conservative force:

i) F1  xiˆ  2yˆj  3zkˆ
 xiˆ  yˆj
ii) F2 
x2  y 2

b) Calculate the work that must be done to lift a satellite of mass 1000 kg to a
height of 1000 km above the surface of the earth. Take the radius of the
earth to be 6400 km and the value of the acceleration due to gravity
g  9.8 ms 2 .

c) A skier of mass M skis 100 m down an incline making an angle  with the
horizontal. On reaching the bottom of the incline she then coasts along the
level ground for 50 m before she comes to a stop. Calculate the coefficient
of friction between the skis and the snow for   30.

Let us now discuss in brief the dynamics of a many-particle system.

1.3 MOTION OF A MANY-PARTICLE SYSTEM


In Sec. 1.2 we studied about the motion of a single particle. Let us now
consider a many-particle system consisting of n particles. Let the i th of these

particles have a mass mi and a position vector ri relative to a given origin,
and i =1,....,n (see Fig. 1.1). So a system of n particles is defined by the set of
 z
masses {mi } and the set of position vectors {ri } , i =1,....,n. Any geometrically 
possible set of values for the position vectors defines a configuration of the r1
system. 
ri

1.3.1 Dynamical Variables and Equations of Motion r2
y
Newton’s second law for the i th particle can be written as:
 dp
Fi  i (1.27) x
dt Fig. 1.1: An n-particle
  system.
Where Fi is the net force acting on the i th particle and pi is its linear
momentum. However now we must distinguish between the two different types
of forces acting on this particle. The first is the net external force acting on the

particle, let us denote that by Fi E . The external force originates from outside
this system of particles. However, because this is a system of many particles,
there are also mutual interaction forces that the particles of the system exert
on one another. These forces originate from inside the system and we call
them the internal forces. So the other type of force acting on the particle is the
sum of all the internal forces
 acting on it due to the other particles in the
system. So we can write Fi as:
  
Fi  Fi E   F ji (1.28)
j i

17
Block 1 The Lagrangian Formulation of Mechanics

Where F ji is the force on the i th particle due to the j th particle. Assuming that
all the masses are constant, we can now write for the i th particle:
   
mi ri  Fi  Fi E  F ji  (1.29)
j i
You would obviously have n such equations, one for each particle. If we wish
to understand the motion of each particle in the system, then we must solve
this set of n second order differential equations. In general these could be
coupled differential equations, for example if the internal forces depend upon
the distance between the i th and j th particles:
k
F ji    (1.30)
n
ri  r j

Let us now see what we can say about the motion of the system of particles
as a whole without solving this set of n differential equations. Summing over
these equations for all the particles in the system, we can write
 E 
 mi ri   Fi   F ji (1.31)
i i i, j
ij

The first term on the right hand side is the sum of all the external forces acting
 
on the system and we denote it by F E   Fi E . Let us examine the second
i
 
term. It contains pairs of terms of the form Fij  F ji , for all values of i and j
( i  j ) . Assuming that these forces obey Newton’s third law, we know that the
force on the ith particle due to the jth particle must be equal and opposite to the
th th
force
 on the j particle due to the i particle, and so for every such pair
Fij  F ji  0 . So we can write:
z  d2  
 mi ri  2  mi ri  F E (1.32)
i dt i

We now define the centre of mass of the system, as the point which is

RCM defined by the following position vector (see Fig. 1.2):
 
y  m i ri  mi ri

RCM  i  i (1.33a)
x  mi M
i
Fig. 1.2: Centre of
mass of an n-particle where M   mi is the total mass of the system. In the Cartesian coordinate
system.  i
system RCM  xCM iˆ  yCM ˆj  zCM kˆ and the expression for the centre of mass
coordinates ( xCM , y CM , zCM ) are
n n
 mi xi  mi xi
xCM  i 1  i 1 (1.33b)
 mi M
n
 mi y i
y CM  i 1 (1.33c)
M

18
Unit 1 Elementary Principles of Mechanics: A Revision
n
 mi z i
zCM  i 1 (1.33d)
M
Using Eq. (1.33a) the equation of motion of the system reduces to:

d 2RCM 
M  FE (1.34a)
2
dt
Which gives us the following set of equations in Cartesian coordinates

d 2 xCM d 2 y CM d 2 zCM
M  FxE ; M  FyE ; M  FzE (1.34b)
2 2 2
dt dt dt
This equation is similar to the equation for a single particle. We can imagine
that a particle whose mass is equal to the total mass of the system of particles,
is located at the centre of mass of the system, and the net external force acts
on it. So as you can see, the motion of the centre of mass of the system
depends only on the net external force acting on the system and purely
internal forces do not influence its motion.
Remember that Eq. (1.34a) tells us only about the motion of the centre of
mass of the system and nothing about the motion/relative motion of the
individual particles of the system. However the concept of the centre of mass
makes its easier to define the other dynamical variables and the total energy
of the system.
We can define the total linear momentum of this many-particle system as the
sum of the linear momentum of each particle of the system and using
Eq. (1.33a) we get

   d   dRCM 
P   mi v i   mi ri    mi ri   M  MvCM (1.35)

dt  i  dt
i i 
So the total linear momentum of the many-particle system is the product
 of z
the total mass of the system and the velocity of its centre of mass vCM which 
ri 
is:
 
  mi v i ri 
 dRCM
vCM   i (1.36) RCM
dt M
y
We can also define the total angular momentum of the system about the
origin in a similar fashion:
 x
   
L   r i  pi   r i  m i v i (1.37) Fig. 1.3: Position
i i vector relative to the
Let us now define the position vector of each particle of the system, relative to centre of mass of the
the centre of mass of the system ( see Fig. 1.3). The position vector of the i th system.

particle relative to the centre of mass , ri  , is:
  
ri   ri  RCM (1.38)
This also gives us the following relation:
  
dri  dri dRCM   
   v i   v i  v CM (1.39)
dt dt dt
19
Block 1 The Lagrangian Formulation of Mechanics

  dri 
Where v i  is the velocity of the i th particle relative to the centre of mass
dt
     
of the system. Replacing r  r   R and v  v   v
i i from Eqs. (1.38)
CM i i CM
and (1.39) respectively in Eq. (1.37), we get

 ri   RCM  mi v i   vCM 


    
L
i

 ri   mi v i   RCM   mi v i    mi ri    vCM


       

i i  i 

  
 RCM   mi v CM
  
 i 

 ri   mi v i   RCM  dt   mi ri      mi ri    vCM


   d       

i  i   i 

  
 RCM   mi v CM
   (1.40)
 i 

Multiplying both sides of Eq. (1.38) by mi and carrying out the summation
over i we get:

   
 mi ri   mi ri  RCM  mi (1.41)
i i i

 
From the definition of the centre of mass we know that  mi ri  MRCM and
i

 mi  M . Substituting in Eq. (1.41) we get  mi ri  0 . Therefore in
i i
Eq. (1.40) the second and third terms would be equal to zero since they

contain  m r  . The final expression for the angular momentum of the n-
i i
i
particle system would be:

 ri   mi v i    RCM    mi vCM


     
L
i  i 

 ri   pi   RCM  P 
   
 (1.42)
i

The total angular momentum of the system about the point O is then the sum
of the angular momentum of total mass of the system located at its centre of
mass , moving with the velocity of the centre of mass of the system and the
angular momentum of each particle of the system about its centre of mass.

20
Unit 1 Elementary Principles of Mechanics: A Revision
SAQ 3

a) Four objects of masses 2.0 kg, 4.0 kg, 1.0 kg and 3.0 kg are located in
the xy-plane, with their position coordinates (in m) being (3.0,2.0),
(1.0,7.0), (1.0, &4.0) and (&2.0, 0) respectively. Determine the centre of
mass of the system.

b) The masses (in kg), position coordinates (in m) and velocities (in ms1) for
a system of four particles is described below:

Mass Position coordinates Velocity


(kg) (m) (ms1)

3.0 (0, 4) 3.0iˆ


1.0 (2, 3) 2.0iˆ  3 ˆj
2.0 (3, 0)  2.0 ĵ
4.0 (3,-2) iˆ  2 ˆj

Obtain the i) the velocity of the centre of mass and ii) the angular momentum
of the system of particles.

Let us now discuss the conservation laws for many-particle systems.


1.3.2 Conservation Laws for Many-Particle Systems
We now study the conditions for the conservation of the total linear
momentum, the total angular momentum and the total mechanical energy of a
many-particle system.
I) Conservation of Total Linear Momentum
Summing Eq. (1.27) over all values of i and using the condition
   
 Fi   Fi E   F ji  F E we get:
i i i, j
i j
 
E dpi dP
F   (1.43)
i dt dt

So if the net external force F E acting
 on the system is zero, then the total
linear momentum of the system, P , is conserved. If the mass of the
system is constant, this also means that in the absence of an external
force, the velocity of the centre of mass of the system is a constant.
II) Conservation of Total Angular Momentum
Let us first calculate the rate of change of the total angular momentum of
the system:

dL d  
dt


dt
ri  pi 
i
  
 dri     dpi    dpi 
  
 pi    ri    ri   (1.44)
i   i 
dt dt  i 
dt 

21
Block 1 The Lagrangian Formulation of Mechanics
 
dr i   
As we know  pi  v i  mi v i  0 . Using Eq. (1.27) and (1.28) in
dt
Eq. (1.44) we get:

     
   ri   Fi E   F ji 
dL
 ri  Fi 
dt  
i i   j i 

  
   
 ri  Fi E  ri  F ji 
i i, j
i j
  
  i   ri  F ji   (1.45)
i i, j
i j
  E
where i  ri  Fi is the torque about the origin, due to external force
acting on the ith particle. So the first term is clearly the net external
 
torque  E   i in the system. The second term can be written as:
i

 ri  F ji    ri  F ji   r j  Fij    ri  r j   F ji 


        
(1.46)
i, j i j i j
ij
 
where we have used the condition Fij  F ji in accordance with Newton’s

z
 
 
third law. Clearly, the vector ri  r j lies along the line joining the ith and jth
 
ri  r j particles (see Fig. 1.4). If the internal force between each pair of particles
also lies along the line joining the two particles, then clearly:
  
ri 
rj 
 

ri  r j  Fij  0 (1.47)

y And the rate of change of the total angular momentum of the many-particle
system is:

x dL  E
Fig. 1.4:  (1.48)
dt
So when the net external torque in the system is zero, the total angular
momentum of the system is conserved.
Most forces in nature do conform to both these conditions which are said
to constitute the strong form of Newton’s third law : that
 is, the action and
reaction forces between two particles i and j, Fij and F ji are not just equal
 

and opposite but also are parallel to ri  r j . For example, the gravitational
force between two particles or the electrostatic force between two
charges. Also note that the conservation of angular momentum for a many
particle system requires that both the conditions are met,
 whereas
 the
conservation of linear momentum required only that Fij  F ji . For the
gravitational force between two particles or the electrostatic force between
two charges, therefore, both the total linear momentum and the total
angular momentum are conserved quantities.
III) Conservation of Total Mechanical Energy
Let us first determine the work done by the all the forces in moving this
many-particle system from a configuration A to a configuration B. Let us

22 say that the configuration A is defined by a set of position vectors {riA } and
Unit 1 Elementary Principles of Mechanics: A Revision

the configuration B is defined by the set of position vectors {riB } . For

example particle 1 would move from the position r1A to the position
   
r1B under the action of the force F1  F1E   F1 j and so on for each
j 1
particle in the system. So the total work done would be:
B   B    
W AB   Fi .dsi     Fi E   F ji  .dsi
i A i A j i 
B  B 
 
  Fi E .dsi    F ji .dsi (1.49)
i A i, j A
i j

where dsi is the infinitesimal displacement of the i th particle along its path of
motion. Let us next relate the work done to the change in the kinetic energy,
when the system moves from configuration A to configuration B.
B   B  B
dv i  1
W AB   Fi .dsi   mi
dt
.v i dt  2
mi d (v i 2 )
i A i A i A

1 1
   2 mi v iB 2  2 mi v iA 2  (1.50)
i

where v iA and v i B are velocities of the ith particle in configuaration A and


1 1
configuration B respectively and TiA  mi v iB 2 and TiB  mi v iB 2 the
2 2
corresponding kinetic energies. So the work done can be related to the
difference in the total kinetic energies in the two configurations as follows:
W AB   TiB   TiA  TB  TA (1.51)
i i

T A and TB are the total kinetic energies of the system in configurations A and
B respectively. This is the work-energy theorem for a system of particles. So
the total kinetic energy for a system of particles in a configuration is:
1
T   Ti   mi v i 2 (1.52)
i i 2
  
Using v i  v i   v CM from Eq. (1.39), we can rewrite the kinetic energy as:

 2 mi v i   vCM 
1   2
T 
i

  2 mi v i  
1  2  1  2  1    
    
  2  mi v CM   2 mi v i .v CM


i  i   i 

(1.53)
As we have already shown :

1  1 d   
 2 mi v i  
2 dt
 m i ri 
  0

(1.54)
i  i 
23
Block 1 The Lagrangian Formulation of Mechanics
The expression for the total kinetic energy reduces to:

  2 mi v i  
1  2 1  2
T    2 Mv CM (1.55)
i 

which is the sum of the total kinetic energy of motion of the particles relative to
the centre of mass of the system and the kinetic energy of the total mass of
the system located at its centre of mass and moving with the velocity of the
centre of mass.
Let us now look at the conditions under which the total energy of a system of
particles will be conserved. As before, this will happen only when all the forces
acting on the system, both external and internal, are conservative. Let us see
what this implies.
 The external forces acting in the system being conservative means that
   
each external force is derivable from a potential. If Fi E   iVi (r1,....., rn )
 
is derivable from the potential energy function Vi  Vi ( r1,....., rn ) , we can
write:
   
Fi E   iVi (r1,....., rn ) (1.56)
 
where  i is the gradient with respect to ri . Therefore the first term on the
RHS of Eq. (1.49) is:
B  B 
     
 Fi E .dsi      iVi (r1,....., rn ) .dsi  Vi riA   Vi riB 
i A i A i i

 VA  VB (1.57)

where VA and VB are respectively the total potential energies of the


system arising out of the external forces, in configurations A and B. In
general, the potential energy function (the potential) arising out of the
external forces acting on the system is:
       
V (r1,....., rn )  V1(r1,....., rn )  V2 (r1,....., rn ).......  Vn (r1,....., rn )


 Vi ({ri }) (1.58)
i

With this the expression for the work done in Eq. (1.49) is:
B  B  B 
  
W AB   Fi E .ds i   F ji .ds i  VA  VB    F ji .dsi
i A i, j A i, j A
i j i j
(1.59)
 For the internal forces to be conservative, let us assume that there is a
potential function Vij , from which the force F ji may be derived as follows:
 
F ji   iVij (1.60)

Let us assume that for a pair of


 particles in the system, i and j, the forces
of mutual interaction Fij and F ji satisfy the following conditions:
24
Unit 1 Elementary Principles of Mechanics: A Revision
 
i) Fij  F ji , and
  
ii) Fij and F ji are parallel to the line joining the two particles, that is Fij
   
and F ji are parallel to the vector rij  ri  r j  
For these conditions to be satisfied we must impose the condition that
the potential function Vij is a function of the distance between the two
  

particles, rij  ri  r j , i.e.

 
Vij  V rij (1.61)

Because Vij is a function only of the distance between the two


particles, Vij  V ji .
  
With this definition, you can see that ( since rij  ri  r j ) :  
  
 
  
F ji   iV rij   jV rij  Fij   (1.62)

So the force of mutual interaction between the particles i and j are


equal and opposite. Further, you can check for yourself:
 
    
 iV rij  h rij rij (1.63)

where h  rij  is some scalar function of


 
rij . Note that this also implies
   
that the force F ji is parallel to rij  ri  r j .  
Therefore, for the internal forces to be conservative (derivable from a
potential function) we must impose, in addition to conditions (i) and (ii),
the following condition:
iii) The magnitude of the internal force must be a function of the distance
  
between the two particles, rij  ri  r j , and  
 
Fij  f ( rij ) rˆij (1.64)

where rˆij is the unit vector along rij .

Now we write the second term in Eq. (1.50) as a summation over pairs
of terms as follows:
B   B   
  
1 
 Fij .dsi 
2 i, j
F ji .dsi  Fij .ds j
i, j A A
ij i j

B  
1
    iVij .dsi   jVij .ds j 
 

2 i, j
A
i j

B  
    iVij .dsi   iVij .ds j 
1  

2 i, j
A
i j

25
Block 1 The Lagrangian Formulation of Mechanics
B 
1  
   iVij . dsi  ds j 
2 i, j
(1.65)
A
i j

The factor of half is added because in rewriting the sum as a sum over
pairs of terms, each term is being counted twice. Now writing
  
  
 iVij   ijV rij   jVij , where  ij represents the gradient of Vij with
  

respect to the variable rij  ri  r j and using 
    
dsi  ds j  dri  dr j  drij , we get

B   B 
1 1 1
  Fij .dri      ijVij .drij   Vij A    Vij B 
i, j A 2 i, j A 2 i, j 2 i, j
ij i j i j i j
(1.66)

where in  Vij A  the summation is over all rij in configuration A and
i, j
i j

in  Vij B  the summation is over all rij in configuration B.
i, j
i j

So the total potential energy function, U, for the system, is the sum of
the potential function arising out of the external forces and that arising
out of the internal forces in the system:
1
U   Vi   Vij (1.67)
i 2 i, j
i j

With this, we can finally write, using Eqs.(1.57) and (1.66):


   
 1   1 
W AB  VA   Vij A   VB   Vij B   U A  UB
 2 i, j   2 i, j 
 i j   i j 
 TB  TA (1.68)
The total energy of the system E  T  U  is conserved. This is the
law of conservation of energy for the many particle system.

SAQ 4
a) Collision (also called scattering) of particles is a very important feature of
the physical universe and much of our knowledge of atoms, molecules,
nuclei and elementary particles has come from scattering experiments.
Consider the elastic collision (a collision in which both linear momentum
and kinetic energy are conserved) between two masses m1 and m2 , in

which initially, m1 is moving with a velocity v and m2 is at rest. After the

collision m1 has a velocity v1 in a direction making an angle 1 with the

original direction of motion and m2 has a velocity v 2 in a direction making
an angle 2 with the original direction of motion. Determine v1, v 2 and 1 .
What is the value of 1   2 when m1  m2 ?

26
Unit 1 Elementary Principles of Mechanics: A Revision
b) Consider the elastic collision of a particle of mass m moving with a

velocity viˆ , with a particle of mass 2m which is initially at rest. If the
kinetic energy of the mass m after collision is one third of its kinetic energy
before the collision, calculate the angles of the direction of motion of the
two particles after the collision.
c) Consider a system of two particles of masses m1 and m2 moving solely
under a force of mutual interaction. Assuming that the force of interaction
depends only on the distance between the two objects, derive the equation
of motion for their relative motion.

With this we complete our brief introduction to the motion of single particle and
many-particles systems studied using Newton’s laws of motion. As you have
seen, all you need is a knowledge of the forces acting in the system and you
can generate the relevant equations of motion, which are second order
differential equations. In principle, you can then determine all the dynamical
variables at each instant of the motion from a set of initial conditions. You can
imagine how this treatment begins to get cumbersome as the number of
particles increases. The simple differential equation for a single particle,
transforms into a set of coupled second order differential equations for many
particle systems.
Notice also that your study of motion depends on your knowledge of forces,
which are vector quantities. To describe vector quantities you need a suitable
coordinate system, which is often dictated by the symmetry of the system you
are studying, for example the two - dimensional Cartesian coordinates for
motion on a plane and plane polar coordinates for planetary motion. The form
of the differential equation shown in Eq. (1.3b) is not the same in other
coordinate systems, as you will study in detail in the next few units.
From the next Unit onwards we study a more abstract formulation of
mechanics, which builds on the principles of Newtonian mechanics but the
description will be in terms of two scalar quantities: the kinetic energy and the
potential energy.

1.4 SUMMARY
 Motion of a Single Particle

In an inertial frame of reference, the equation of motion for a particle


 dp d mv 
of mass m is written using Newton’s second law as: F  
dt dt
 
The linear momentum is p  mv .For constant mass, the equation of
motion is:
 
dv  
F m  ma  mr
dt

In the Cartesian coordinate system, with r (t )  x(t )iˆ  y (t ) ˆj  z(t )kˆ , we
get the following set of three differential equations:
mx  Fx ; my  Fy ; mz  Fz
27
Block 1 The Lagrangian Formulation of Mechanics

If the position of the particle with respect to a point O is given
 by r ,
  
then the angular momentum L of the particle about O is: L  r  p

The torque  acting on the particle, also called the moment of the

     dL
force F , about the point O is :   r  F and   .
dt

 Conservation Laws for a single particle system

i) If the total force acting on a particle is zero, the linear momentum is


conserved.
ii) If the total torque acting on a particle is zero, then the angular
momentum is conserved.
iii) If the forces acting on a particle are conservative, then the total
mechanical energy of the system is conserved.

 Motion of a Many-Particle System


A system of n particles is defined by the set of masses {mi } and the set

of position vectors {ri } , i =1,....,n. The equation of motion for the ith
particle is
  
mi ri  Fi E   F ji
j i
 
Fi E is the net external force acting on the particle and F ji is the force
on the ith particle due to the jth particle.

The centre of mass of the system is defined by the position vector


 
  mi ri  mi ri
RCM  i  i
 mi M
i

The motion of the centre of mass of the system depends only on the
net external force acting on the system and the equation of motion is:

d 2RCM 
M  FE
2
dt

The total linear momentum of the system is the product of the


 total
mass of the system and the velocity of its centre of mass vCM
  
P   mi v i  Mv CM
i

The total angular momentum about a point O is the sum of the angular
momentum of total mass of the system located at its centre of mass ,
moving with the velocity of the centre of mass of the system and the
angular momentum of each particle of the system about its centre of
mass:
     
 
 

L   ri  pi   ri  mi v i    ri   pi   RCM  P
i i i
28
Unit 1 Elementary Principles of Mechanics: A Revision
  
where ri   ri  RCM .

 Conservation Laws for Many-Particle Systems



i) If the net external force F E acting on the

system is zero, then the
total linear momentum of the system, P , is conserved.
 
ii) If the forces of mutual interaction Fij and F ji satisfy the following
conditions:
 
 Fij  F ji , and
 
 Fij and F ji are parallel to the line joining the two particles, that is
  

 
Fij and F ji are parallel to the vector rij  ri  r j 
then the rate of change
 of the total angular momentum of the many-
dL  E
particle system is   and when the net external torque in the
dt
system is zero, the total angular momentum of the system is
conserved.
iii) In addition, when all the forces acting on the system, both external
   
and internal are conservative ( Fi E   iVi (r1,....., rn ) and
 
Fij  f ( rij ) rˆij ) ), then total energy of the system E  T  U  is
conserved.

1.5 TERMINAL QUESTIONS


1. A box of mass 1 kg is held in place on a fixed inclined plane by the force of
static friction between box and the plane. If the coefficient of static friction
between the box and the plane is 0.3, determine the angle between the
box and the plane at which the box will begin to slide down the plane.
2. A projectile of mass m is launched from the ground with a speed v0 in a
direction which makes an angle  with the horizontal. If the particle

encounters a force of resistance  mkv , determine the projectile’s position
and velocity.
3. Two masses M and 3M are suspended on either side of a smooth pulley at
rest by a massless inextensible string as shown in Fig. 1.5 (Atwood’s
machine). Calculate the acceleration of each mass and the tension in the
string. What would be the acceleration of each mass if the pulley is placed
in an elevator moving downwards with a constant acceleration .
M
4. A ballistic pendulum is a device used to measure the speed of a bullet. The
pendulum is a large wooden block of mass M hanging vertically. A bullet of

mass m strikes the pendulum with an initial velocity u and gets embedded 3M
in it. The impact of the bullet causes the pendulum to rise to a height h.

Calculate u . Fig. 1.5: Atwood’s
5. A block of mass M sliding along a horizontal table with speed u, hits a Machine.
spring fixed to the wall, which has a spring constant k. The left end of the
spring is initially at x =0. On hitting the spring it experiences a frictional
force with a variable coefficient of friction given by   ax , where a is a
29
Block 1 The Lagrangian Formulation of Mechanics
constant. Calculate the loss in mechanical energy and the compression in
the string when the block first comes to rest.
6. A meteor of mass 10.0 kg, enters the Earth’s atmosphere with an initial
 
velocity of V0  3.0 km s 1 ˆj and an acceleration of a  2.0 m s 2 ˆj . A
time t after it enters the atmosphere, it is observed by someone on the
ground. At the moment it is observed, its position relative to the observer is

r  20.0 km iˆ  25.0 km ˆj . Calculate a) the angular momentum of the
meteor about the origin, located at the observer? b) the torque on the
meteor about the origin.
7. If the centre of mass of three objects in the xy-plane is given by the
location (3 cm, 4 cm). Two of the objects have the following mass and
locations: 5 kg at (3 cm,2 cm), 2 kg at (1 cm, 3 cm) and the third object
has mass 8 kg. Determine the location of the third object.
8. A projectile of mass M is launched from the ground explodes into three

fragments A, B and C at some point in its flight, when it’s velocity is v 0 .
One fragment, which has a mass M A  M / 3 continues in the original
direction of the projectile at the moment of explosion. The second fragment
which has a mass M B  M / 6 , moves in the opposite direction and the
third fragment which has a mass MC  M / 2 comes to rest. The energy
released in this explosion is five times the original kinetic energy. Calculate
the speed of the fragments A and B.
9. Consider the head on elastic collision between two particles of masses m1
and m2 which are both travelling in the same direction with velocities
 
u1 and u2 respectively. If the kinetic energy of the two particles is equal
before the collision, calculate the values of m1 / m2 and u1 / u 2 for which
the mass m2 will be at rest after the collision.
10. A proton (mass mp) travelling at 2.0  10 6 m/s undergoes an elastic
collision with an alpha particle which is initially at rest. After the collision
the proton is deflected at an angle of 60° with respect to its initial direction
of motion. Given that the mass of the alpha particle is 4m p , calculate what
percent of its initial kinetic energy the proton would retain after the
collision.

y 1.6 SOLUTIONS AND ANSWERS



N
Self-Assessment Questions
1. a) The
 forces acting on the box are the gravitationalforce
x Fg  Mg sin  iˆ  Mg cos ˆj and the normal force N  Nˆj using the
Mg sin 
 Mg cos  coordinate system shown in Fig. 1.6. Writing Newton’s second law for
Fg the system in the component form (Eq. 1.3b), we have

Fx  Mx  Mg sin  (i)
Fig. 1.6: Box on an
inclined plane
Fy  My  N  Mg cos 

Since there is no motion along the y-direction we get y  Fy  0 and


N  mg cos  . We therefore have to solve the ODE in Eq. (i) which is
30
Unit 1 Elementary Principles of Mechanics: A Revision
 x  g sin   x  gt sin   v 0

t2
and x  g sin   v 0t  x0
2
where v 0 and x0 are constants to be determined from the initial
conditions . At t  0 , x  0  v 0  0 and x  0  x0  0 . So
t2
x  g sin  .
2
b) We write down the equation of motion assuming the motion is along
the
 z-direction (Fig. 1.7)and the forces
 are the gravitational force
ˆ
Fg  mgk and the retarding force FR  av z kˆ  av z kˆ , where a is a
constant. The equation of motion is
dv z
m  mg  av z
dt
  
at

mg 
which has the solution v z  1  e m  where we have assumed FR
a  
 
dz 
that v z  0 at t = 0. Writing v z  we get the differential equation: Fg
dt z
  
at
dz mg  
 1 e m
dt a  
Fig. 1.7
 
and on solving, With z  0 at t = 0, we get:
at
mgt  m 2g  
z   1 e m
a  a2 
 
c) Using Eq.(1.7) we write the angular momentum
    
L  r  p  r  (m v )
 ( 4.0 m iˆ  2.0 m ˆj )  3.0 kg (5.0 ms 1 iˆ  3.0 ms 1 ˆj )

 6.0 kg m2s 1 k̂
The torque (Eq. 1.8) is
   B
  r  F  ( 4.0 m iˆ  2.0 m jˆ )  (8.0 N jˆ)  32 N m k̂
   h
2. a) For a conservative force field F ,   F  0
     A
i)   F1    (2 xiˆ  2yˆj  3zkˆ )  0 . Hence F1is conservative
   ( x iˆ  y ˆj ) 4 xy RE
ii)   F2     kˆ
2
x y 2
(x  y 2 )
2
 rˆ
Hence F2 is not conservative . O

b) The force F on the satellite (mass Ms) is the gravitational force:
  GM M
F E s rˆ
r2
Fig. 1.8
where ME is the mass of the earth Ms is the mass of the satellite and r
is distance of the satellite from the centre of the earth (Fig.1.8). The
work done in raising the satellite from a point A on the surface of the 31
Block 1 The Lagrangian Formulation of Mechanics
earth to B at a height h  1000 km  10 6 m from the surface of the earth
is:
B  B GM M rˆ
W AB   F . dr    E s .dr rˆ
A A r2

(dr  dr rˆ )
B r
dr  GME Ms  B 1 1
 G ME Ms  2     GME Ms   
 r  rA r
 B r A
Ar

Using r A  RE  6400 km  6.4  10 6 m ,


rB  RE  h  7400 km  7.4  10 6 m , G  6.67  10 11 Nm 2kg2 ,
ME  6.0  10 24 kg and M s  10 3 kg , the magnitude of the work
needed is 8.4  109 J .

c) 
N1

N2 
d1 Ffr
 h
 Fg
V 
d2 
Fg
Fig. 1.9

In this problem, we have a non conservative force, which is friction.


The change in the mechanical energy is the work done against the
average frictional force. The forces acting on the skier
 on the incline,
are the gravitational forces Fg , the normal force N1 and the frictional

force Ffr , with Ffr  s N1 . We know that N1  Mg cos  and so
Ffr   s Mg cos  .

The change in the mechanical energy when the skier gets to the
bottom of the incline, is the work done against friction over the incline
(Fig. 1.9)
1
 Mg h  MV 2  S Mg cos  d1
2
V is the speed of the skier at the bottom of the incline and d1  100 m is
the length of the incline and h  d1 sin  is the height of the incline. So
1
Mg d1 sin   s Mg cos d1  MV 2 (i)
2
On reaching bottom of the incline, the skier starting with a speed of V,
moves on level ground till it come to rest. The normal force is now
N2  Mg . The law of conservation of energy gives us:
1
MV 2  s Mg d 2 (ii)
2
where d 2  50 m is the distance covered by the skier before she comes
32 to rest. Using Eq. (ii) in Eq. (i) we get:
Unit 1 Elementary Principles of Mechanics: A Revision
d1 sin    s cos d1   s d 2
d1 sin  100 sin 30
 s    0.36
(d1 cos   d 2 ) 100 cos 30  50

3. a) Using Eqs. (1.33b-d) of SAQ 3a, we can write:


2.0 kg 3.0 m 4.0 kg 1.0 m 1.0 kg 1.0 m 3.0 kg ( 2.0 m)
xCM 
2.0 kg 4.0 kg 1.0 kg 3.0 kg
 0 .5 m
2.0 kg 2.0 m 4.0 kg 7.0 m  1.0 kg 4.0 m 3.0 kg 0
yCM   2.8 m
10.0 kg
b) Using Eq. (1.36) the velocity of the centre of mass is
4
 mi v i
v CM  i 1
M


(3.0  3.0 i )  1.0  (2.0 iˆ  3.0 j )  2.0  2.0 jˆ)  4.0  (iˆ  2 ˆj )kg ms1
10 kg
1 ˆ 1
 [9i  2iˆ  3 ˆj  4 ˆj  4iˆ  8 ˆj ] ms 1  [15iˆ  ˆj ] ms 1
10 10
Using Eq. (1.37) the angular momentum of the system is
  

L   ri  pi  ( 4 ˆj )  3.0(3.0iˆ)  (2iˆ  3 jˆ)  1.0(2.0iˆ  3.0 ˆj )
i

 ( 3.0iˆ)  2.0( 2.0 jˆ)  (3iˆ  2 jˆ)  4(iˆ  2 ˆj ) kgm 2 s 1

 4 kgm 2 s 1 k̂
y 
4. a) m1 v1


v 1
x
m1 m2 2
Before collision m2

(a) v2
After collision
(b)
Fig. 1.10: Elastic collision between two particles in two-dimensions.
 
In Fig. 1.10b, 1 and 2 are the angles that v1 and v 2 make with the
x-axis, respectively. Equating the components of the linear momentum
before and after the collision(conservation of linear momentum) we get
m1v  m1v1 cos 1  m2v 2 cos 2 (i)

and 0  m1v1 sin 1  m2v 2 sin 2 (ii)

Since the kinetic energy of the system remains constant, we have

m1v 2  m1v 12  m 2v 22 (iii)

To eliminate 1 from Eqs. (i and ii), we can rewrite them as


33
Block 1 The Lagrangian Formulation of Mechanics
m1v1 cos 1  m1v  m2v 2 cos 2 (iv)

and m1v1 sin 1  m2v 2 sin 2 (v)

Squaring Eqs. (iv and v) and adding the resulting equations, we get:

(m1v1)2 (cos 2 1  sin2 1)

 (m1v  m2v 2 cos 2 )2  (m2v 2 sin 2 )2

or m12v12  m22v 22  m12v 2  2m1m2v v 2 cos 2 (vi)

Multiplying Eq. (iii) by m1 , we get

m12v 2  m12v12  m1m2v 22 (vii)

Replacing Eq. (vii) in Eq.(vi) we get

0  m22v 22  m1m2v 22  2m1m2v v 2 cos 2 (viii)

Eq. (viii) is a quadratic equation in v 2 of which v 2  0 is a trivial


solution. We disregard that and write only the acceptable solution,
which is
2m1v cos 2 2v cos  2 m
v2   where   1 (ix)
m1  m2 1  m2

We can determine the value of v1 from Eq. (i).


If m1  m2 ,   1 and from Eq. (ix) we get that v 2  v cos 2 .
Therefore, from Eqs. (iv) and (v) we get
sin 22 2 sin 2 cos 2
tan 1    cot 2  tan(90  2 )
1  cos 22 2 sin2 2

 1  90  2 or 1  2  90

b) The kinetic energy of the mass m before and after the collision be E
and E1 respectively and that of the mass 2m after the collision be E2.
The velocity and angle with the original direction of motion of the mass
 
m and 2m after collision are v 1 ,1 and v 2 ,2 respectively (refer SAQ
4a).
E 2E
 E  E1  E2 and E1   E2 
3 3
We have
1 1 1 1  1 2 1 
E  mv 2 ; E1  mv12   mv 2  ; E3  (2m )v 22   mv 2 
2 2 32  2 3  2 
1 2 2 1 2 1
So v12  v ; v 2  v  v1  v 2  v
3 3 3
m 1
Using Eq. (ix) of SAQ 4(a) with    we get:
2m 2
3
cos 2    2  30 
34 2
Unit 1 Elementary Principles of Mechanics: A Revision
And using Eq. (v) of SAQ 4(a) we get sin 1  1  1  90
  
c) rˆ12 is the unit vector along r12  r1  r2 (Fig. 1.11). The force acting on m1   
    r12  r1  r2
m1 is along (r1  r2 ) and depends only on r12  r1  r2 , so we can write m2
 
 r1
F1  F1( r12 )rˆ12  m1r1 (i) 
  r2
Since it is a force of mutual interaction, F1  F2
  O
F2  F1(r12 ) rˆ12  m2 r2 (ii)
Fig. 1.11
Multiplying Eq. (i) by m2 and Eq.(ii) by m1 we get

m1m2r1  m2F1(r12 ) rˆ12
and
m1m2 r2  m1 F1( r12 ) rˆ12
  m  m2
 r1  r2  1 F1(r12 ) rˆ12
m1m2
 1 m1m2
or r12  F1(r12 ) rˆ12 where   r12
 m1  m2

  r12  F1(r12 ) rˆ12
This is the equation of motion for relative motion.
Terminal Questions
1. The box will begin to move when the downward  force overcomes the force
y
of static friction. The gravitational force is Fg  mg cos  ˆj  mg sin  iˆ, the
 
normal force is N  Nˆj and the frictional force is F   Niˆ (Fig. 1.12). So:
fr s 
N
Fx  mg sin    s N ; Fy  0  mg cos   N  N  mg cos   x
Ffr
 Fx  mg sin    s mg cos 
Mg sin 
The box will move when Fx  0 , that is  Mg cos 
Fg
mg sin   s mg cos   tan   0.3 for s  0.3 
or   tan 1 0.3    16.7
Fig. 1.12
2. The initial velocity of the projectile(Fig. 1.13) is

v  v cos  iˆ  v sin  jˆ
0 0 0 (i)

The position of the particle at time t is r  x iˆ  y ˆj y

Forces acting on the projectile are: 


 v
 FR
 the force of gravitation Fg  mg ˆj and 
  Fg
 v0
 the force of resistance FR  mk v
 
writing v  x iˆ  y ˆj we get x
 Fig. 1.13
FR  mk x iˆ  mk y ˆj (ii)
The equations of motion are:
Fx   mx  mk x  x  k x  0 (iii)

Fy  my  mg  mk y  y  k y  g (iv)
35
Block 1 The Lagrangian Formulation of Mechanics
On solving and using the initial conditions:
x (t  0)  y (t  0)  0 and x (t  0)  v 0 cos ; y (t  0)  v 0 sin 

We get:
v cos  gt kv sin   g
x (t )  0 (1  e  kt ) ; y (t )    0 2 (1  e  kt )
k k k

O 3. There is no friction between the string and the pulley and tension T is
same throughout the string. For the two masses m1  M and m2  3M we
can write the equations of motion as (Fig. 1.14a):
y1 m1y1  m1g  T  T  m1g  m1y1 (i)
y2
m1 m2 y2  m2g  T  T  m2g  m2 y2 (ii)
y Since the string in inextensible we must have y1   y2 . So
m2 m1gˆj m1y1  m1g  (m2 g  m2 y2 ) (iii)

m2 y2  m2 g  (m1g  m1y1 ) (iv)

Using y1   y2, in Eq. (iii) we get


m2gˆj g (m1  m2 )
mi y1  m1g  m2g  m2 y1  y1   y2  (v)
(a) m1  m2

The tension in the string is (from Eq. i)


y 2 y1 g (m1  m2 )m1 2m1m2g
y 2 y1 T  m1g  m1y1  mi g  
m1  m2 m1  m2

For m1  M and m2  3M
O  2Mg g g 3
y1   y2    y2  and T  Mg .
4M 2 2 2
y1
y2 When the pulley is in the elevator, the coordinate system located at the
centre of the pulley is not an inertial frame, since the pulley is accelerating.
m1 We choose an inertial frame which has its origin at the top of the elevator
shaft. Now the coordinate of the masses are y1  y1  y1 and y 2  y1  y 2 .
y
The equations of motion are:
m2 m1gˆj
m1y1  m1g  T  m1y1  m1y1 (v)

m1y2  m2g  T  m2 y2  m2 y2 (vi)


Now, y1  y2   and y1   y2 , so
m2gˆj
m1y1  m1g  T  m1  m1(g   )  T (vii)
(b)
And m2 y2  m2 ( g   )  T (viii)
Fig. 1.14: a) Atwood’s
Machine with pulley at rest Solving we get for m1  M and m2  3M
b) in an accelerating
elevator. (m1  m2 ) 1
y1   y2  (g   )   (g   )
m1  m2 2

2m1m2 (g   ) 3
and T   M (g   )
m1  m2 2
36
Unit 1 Elementary Principles of Mechanics: A Revision
The results for the acceleration and tension are just as if the motion is with an
effective acceleration due to gravity g   g   .

4. The initial linear momentum of the bullet is m u . The linear momentum of


the system after collision is ( m  M ) v f .So from conservation of linear momentum,

mu  (m  M ) v f
1
The kinetic energy before collision is mu 2 and the total kinetic energy after
2
1 1 m 2u 2
collision is (m  M )v f2  . As you can see there is a loss of kinetic
2 2 mM
energy. Hence, the collision is inelastic. However, the kinetic energy after M
h
collision makes the wooden block swing up to a maximum height, h ,as shown in M

Fig.1.15. The kinetic energy of the bullet and block is used up in raising the block u
through a height h . So the block and bullet acquire a potential energy equal to
Fig. 1.15 Ballistic
(M  m ) gh . The kinetic energy of the block and bullet Pendulum
1
is (M  m )v f2  (M  m ) gh, from the principle of conservation of energy. So
2
M m
v f2  2gh or v f  2gh . Hence from Eq. (i), we get u  2gh .
m

1
5. The initial energy is Ei  Mu 2 (Fig. 1.16). Let us assume that the block
2
comes to rest when the spring is compressed by d. So the final energy is
1 1 1
E f  k d 2 . The loss in mechanical energy is E f  E i  kd 2  Mu 2 .
2 2 2
The work done on the system (Fig. 1.16) for is   ax is:

 E f  E i  Work done on the System



u
d d 1
W    Ndx    ax Mg dx   Mgad 2 M
0 0 2
x 0 x
The work done on the system is equal to E f  E i . So we get
Fig. 1.16
1 2 1 1
kd  Mu 2   Mgad 2
2 2 2

M
 d u
k  Mg a

6. Velocity
 at a time t after the meteor enters the earth’s atmosphere is
x V  Vx iˆ  Vy ˆj (Fig. 1.16). Given that the initial velocity is
 
V0  3.0 km s 1 ˆj and a  2.0 ms 2 ˆj , we have a x  0 and

 
a y  2.0 ms 2 . So V   3.0  10 3  2.0 t ms 1 ˆj . Using the equation for
  
the angular momentum L  r  mv with m  10.0 kg and

r  2.0  10 4 m iˆ  2.5  10 4 m ˆj , we get

37
Block 1 The Lagrangian Formulation of Mechanics


 
L  (2.0  10 4 iˆ  2.5  10 4 ˆj )  10( 3.0  10 3 ˆj  2.0t ˆj ) kgm 2 s 1

 (6.0  10 8  4.0  10 5 t ) kg m2 s 1 kˆ

 dL
The torque is:    4.0  10 5 Nm kˆ
dt

7. Let the coordinates of the 8 kg mass be (x,y). So,

5 kg 3 cm 2 kg 1cm 8 kg x 7


xCM  3 cm   x  cm
15 kg 2

5 kg 2 cm 2 kg 3 cm 8 kg y 11


y CM  4 cm   y  cm .
15 kg 2

8. y
 y 
u VA

VB MA
M
MB
MC
x x

Just before the explosion After the explosion

Fig. 1.17

The velocity of the projectile at the point it explodes is V  uuˆ (Fig 1.17).
M M
After explosion the masses of fragments are M A  , MB  and
3 6
M  
MC  . Their velocities after the explosion are VA  V Auˆ, VB  VB uˆ
2
and VC  0 respectively. From the conservation of linear momentum we
can write:
   
M A V  M AVA  M BVB  MCVC  Muuˆ  M AVAuˆ  M BVB uˆ

 Mu  M AV A  M BVB

V V
 u A  B (i)
3 6

1 
If E  5 Mu 2  is the energy released in the explosion, from the
2 
conservation of energy we get
2
1 1 1 V 2 V
E  Mu 2  M AVA 2  M BVB2  6u 2  A  B (ii)
2 2 2 3 6

Using Eq. (i) in Eq. (ii), we can write

38
Unit 1 Elementary Principles of Mechanics: A Revision
2
V2 V 
6u 2  A  6 A  u   VA  4u and VB  2u .
3  3 
  
9. u1 u2 v1

m1 m1
m2 m2

Before After
Fig. 1.18

Since the initial kinetic energies are equal, we have (Fig. 1.18)
1 1
m1u12  m2u22  m1u12  m2u22
2 2
m1
Let   2   2u12  u 22  u1  u 2
m2
From the conservation of linear momentum,
m1u1  m2u2  m1v1  m1u1  m2u1  m1v1
 (m1  m 2 )u1  m1v 1
2     1
And ( 2   )u1   2 v1  v 1  u1    u1
2   
From the conservation of kinetic energy we get
1 1 1 1
m1u12  m2u 22  m1v 12  m1u12  m1v 12
2 2 2 2
2
1 2 1    1 2
or u12  v1    u1
2 2  
2
   1 2 2
    2    1  2  2
  
2 44
This gives us the equation  2  2  1  0     1 2
2
and   2.414
m1 u2
  (2.414 ) 2  5.83 and  2.414 .
m2 u1
10. y

v1


u 60  x

mp 4m p

v2

Before After

Fig. 1.19

From the conservation of linear momentum we can write (Fig. 1.19)

mpu  mpv1 cos 60  4mpv 2 cos  (i)

and 0  mpv1 sin 60  4mpv 2 sin  (ii)


39
Block 1 The Lagrangian Formulation of Mechanics
From the conservation of energy we get

1 1 1
m pu 2  m pv 12  ( 4m p )v 22  4v 22  u 2  v12 (iii)
2 2 2

From (i) & (ii) we get

4v 2 cos   u  v 1 cos 60 (iv)

and 4v 2 sin   v1 sin 60 (v)

Squaring and adding Eqs. (iv) and (v) we get

16v 22 (cos 2   sin2 )  u 2  v 12 (sin2 60   cos 2 60  )  2 u v1 cos 60 

 16v 22  u 2  v 12  uv 1 (vi)

Using (iii) in (vi) we get

4(u 2  v 12 )  u 2  v 12  uv 1 or 5v 12  uv 1  3u 2  0 (vii)

 1  1  60 
Solving Eq. (vii) we get v1    u  0.88u

 10 

1
m p 0.88u 2
And the ratio of the kinetic energies is: 2  77% .
1 2
mp u
2

40
Unit 2 Constrained Motion and the D’Alembert’s Principle

UNIT 2
CONSTRAINED MOTION
AND THE D'ALEMBERT’S
PRINCIPLE
Structure

2.1 Introduction 2.3 D'Alembert’s Principle


Expected Learning Outcomes Virtual Work
2.2 Constraints D'Alembert’s Principle
Examples of Constrained Motion 2.4 Generalized Coordinates
Classification of Constraints 2.5 Summary
2.6 Terminal Questions
2.7 Solutions and Answers

2.1 INTRODUCTION
In Unit 1, we revised the basic concepts of Newtonian Mechanics as applied to
the motion of a single particle, as well as to a many-particle system. Often the
study of mechanics is classified into three broad areas, the Newtonian
formulation, the Lagrangian formulation and the Hamiltonian Formulation. In
this Unit we study the foundations of the Lagrangian formulation of classical
mechanics, which is a whole new way of solving problems in mechanics.
You saw in Unit 1, that central to the application of Newton’s second law in
writing down the differential equation governing the motion of the system, is a
knowledge of the forces acting upon the system. The position of a particle is
completely defined by its position vector, which translates to three position
coordinates in the Cartesian coordinate system. For an n-particle system, you
get 3n second order differential equations and on solving these differential
equation you know everything that can be known about the motion. However
this method could become tedious in certain situations, for example when the
motion of the particle or system of particles in constrained in some way. We
discuss the notion of constraints in motion in Sec. 2.2. Constrained motion
forces us to revisit the notion of degrees of freedom and the number of
coordinates required to study the motion of a system. In Sec. 2.3 we first
introduce the idea of virtual work and the principle of virtual work to define
the condition of equilibrium of a system. The Euler-Lagrange equations of
motion, which we shall discuss in Unit 3, are derived from a fundamental result
of classical mechanics, the D'Alembert ’s principle, sometimes called the
Lagrange-D'Alembert principle, developed by D'Alembert in 1743 to solve 41
Block 1 The Lagrangian Formulation of Mechanics
dynamical problems. D'Alembert’s principle is a statement of the fundamental
laws of motion and we discuss this next in Sec. 2.3. We introduce the
concept of generalized coordinates in Sec. 2.4.
Expected Learning Outcomes
After studying this unit, you should be able to:
 identify and classify the constraints in a dynamical system;
 determine the number of degrees of freedom in a dynamical system;
 state and derive the principle of virtual work;
 apply the principle of virtual work to obtain the condition of static
equilibrium and acceleration of a constrained dynamical system;
 state and derive the D'Alembert’s principle;
 determine a set of generalized coordinates for a constrained dynamical
system; and
 apply D’Alembert’s principle to constrained dynamical systems.

2.2 CONSTRAINTS
We discussed the idea of degrees of freedom in Unit 1. Remember that for
the unconstrained motion of a particle, which essentially means that the
particles can move anywhere in (three-dimensional) space, the number of
(dynamical) degrees of freedom are 3 for a single particle and 3N for an N-
particle system respectively. So the position of the particle is described by 3
coordinates and for an N-particle system the configuration is completely
specified by 3N coordinates.
But what if the motion of the system is constrained in some way? Suppose
the particle or system of particles is restricted to move in a particular way,
such that the coordinates of the particle are required to satisfy certain given
conditions? The prescribed path or surface along which motion takes place is
called a constraint and the equations that describe the path or surface along
which motion is constrained to take place are called the constraint equations.
That would imply that the coordinates for each particle may not vary
independently of each other. There are several dynamical systems in which
this is true. There are offcourse certain forces that are responsible for creating
these constraints on the motion of the particle/particles in the system. These
forces are called the forces of constraint. For example:
 A box sliding down an inclined plane is constrained to slide along the
surface of the plane, and for that the surface of the plane exerts a contact
force(the normal force) on the box at all times.
 When a gas is confined in a container, the walls of the container exert a
force on the molecules to keep them confined within the container.
To use Newtonian mechanics we have to incorporate the forces of constraint
along with the other forces into our equations of motion and then find a
solution to the dynamical problem. In general, the equation of motion for a
single particle undergoing constrained motion can be written as:
42
Unit 2 Constrained Motion and the D’Alembert’s Principle
  
mr  F E  F C (2.1)
C
where F is the force of constraint. Some typical constraint forces that you are
familiar with are the contact forces, tension in an ideal string and the force of
static friction ( when the object is not slipping). Typically, unlike the external
forces, the constraint forces are not explicilty known to us and do not usually
follow a force law. Therefore Eq. (2.1) is a set of three scalar differential
equations for x(t), y(t), z(t) (in the Cartesian coordinate system though you
could also choose any other set of three coordinates) and six unknowns,
which are the position coordinates and the three components of the constraint
force.
Clearly you cannot arrive at a unique solution for the motion of the particle
from Eq. (2.1) unless you know the forces of constraint explicitly.
Let us consider the following examples of constrained motion, which you all
familiar with.
2.2.1 Examples of Constrained Motion
a) Motion of a Simple Pendulum
The position coordinates of the bob at any instant of time (Fig. 2.1) are
( x, y , z ) . Since the simple pendulum moves in a plane, we know that the
z coordinate of the bob is always zero:
z0 (2.2a)
Further the motion of the bob is also restricted by the condition that the
bob is tied to the support by string of constant length L0, hence:
z
2 2 2
x  y  L0 (2.2b) O
x
Eq. (2.2a and b) are the equations defining the constraints in the system.
 L0
x and y are no longer independent variables because: y

x  L02  y 2 (2.3a)
x
And any variation of x, implies a corresponding variation in y. To locate the
the position of the pendulum in the plane of its motion we need just one y
dynamical variable, which could be either x or y or even some other
Fig. 2.1: Simple
variable. For example, you can see that the x and y coordinates are
Pendulum.
related to the angle  as follows:
x  L0 sin  ; y  L0 cos  (2.3b)
So the position of the pendulum can also be uniquely described by just
the one angular variable . What is important is that, because of the two
constraints in the system (Eq. 2.2a and b ), the number of independent
dynamical variables in the system is reduced from three to one and we
say that the pendulum has just one dynamical degree of freedom.
In general constraints in the system reduce the number of dynamical
degrees of freedom of the system.
The force of constraint in the system is the tension in the string. You may
remember that whenever you have studied the motion of the simple
43
Block 1 The Lagrangian Formulation of Mechanics
pendulum in mechanics, you would have derived the expression for the
tension in the string to solve the problem.
Let us now look at another simple example of constrained motion.
b) Atwood Machine
z
The Atwood machine, consists of two blocks of wood (Fig. 2.2) connected
O by an inextensible string. The coordinates of the two blocks are x1, y1, z1 
x and x2 , y 2 , z2  respectively. However since the motion of both blocks is
effectively one dimensional ( x1  x 2  z1  z 2  0 ) let us say that we
y2
describe the motion using two dynamical variables, y1 and y2. So the
y1
system has two degrees of freedom. However because both the blocks are
m2 connected by a fixed length of string, any increase in y1, means an equal
y decrease in y2. Effectively, assuming that the pulley has zero radius, we
m1 can write:

y1  y 2  l (2.4)
Fig. 2.2: Atwood’s
Machine. where l is the length of the string joining the two blocks. This is the
constraint equation. So now, we can describe the motion of the system
z using just one variable either y1 or y 2 (Fig. 2.2) and the system has just
one degree of freedom. This is another example of a constraint, which
reduces the number of degrees of freedom.
 
 ri  r j
ri Let us look at another example where the effect of constraints is striking.
 c) Rigid Body
rj
A rigid body (Fig 2.3) is a system of n particles in which each particle is at
y a fixed distance from every other particle. For every pair of particles, we
can write:
x  
Fig. 2.3: Rigid
ri  r j  c ij , i , j  1,2,...., n (2.5)
Body.
where c ij  c ji  0 . However an unconstrained rigid body, irrespective of
the number of particles in it, can have at most 6 degrees of freedom, and
z
not 3n, which is a result of the constraint contained Eq. (2.5). You will
understand how that is so when you study rigid body dynamics in your
O x
A
next semester course Classical Mechanics-II.
LA
yA There are several other dynamical systems in which the motion is constrained
yB in some way.
xA
A Example 2.1
B LB
Consider a system consisting of two simple pendulums connected in series
B (Fig. 2.4). A system of this kind is also called a double pendulum. Write
xB down the equations of constraint and determine the number of degrees of
freedom of the system.
y
Solution : To begin with, we have two bobs A and B, with coordinates
Fig. 2.4: Double x A, y A, zA  and xB , y B , zB  respectively, choosing the point of suspension O
Pendulum. to be the origin of the coordinate system. Since the pendulums move in a
plane:
44
Unit 2 Constrained Motion and the D’Alembert’s Principle
z A  zB  0 (i)

The motion of the particles is further constrained by the (inextensible) strings


of lengths L A and LB , connecting A to O and B to A, respectively, as shown
in the figure. The equations of constraint are:

x A 2  y A 2  LA 2 (ii)

And
x B  x A 2  y B  y A 2  LB 2 (iii)

We have four constraints in the system (Eqs. i, ii and iii). So we need to use
just two coordinates to locate A and B in the xy-plane. These could be either
the  x A , x B  or y A , y B  . As in the case of the simple pendulum, we could
also choose the angles  A and B , which are the angular displacement of the
masses relative to the vertical, to describe the motion(as shown in Fig. 2.4).

SAQ 1
Write down the equation(s) of constraint for the following systems:
a) A tiny particle of mass m sliding without friction down the surface of a
sphere of radius R under the action of gravity.
b) A simple pendulum of mass m, and length L0 , suspended from a point of
support of mass M which is sliding horizontally.
c) A bead sliding due to gravity on a fixed elliptical wire.
d) A spherical pendulum consisting of a mass m fixed to the end of light
inextensible string of length L0 where the mass is free to move in any
direction (as long as the string remains taut).

Sometimes, inspite of constraints, the dynamics of the system can easily be


determind using the Newtonian formulation, but not always. We now move
towards a more formal description of constraints within a dynamical system as
a first step towards a different formulation of mechanics in which this problem
can be addressed.
2.2.2 Classification of Constraints
The conditions defining the constraints in a system are typically described in
terms of relationships between the dynamical variables of the system and
time. Based on the nature of these relationships, the constraints may be
classified in the following different ways:
i) Holonomic and Nonholonomic Constraints
Consider an N-particle system described by the set of position vectors
  
r1, r2 ,......, rN . 3N coordinates are required to describe the unconstrained
motion of this system. Suppose now that there are constraints in the
system. If a constraint condition (there may be more than one, say k ) is
defined by an equation connecting the coordinates of the particles (and
possibly time), of the following form:
   45
f ( r1, r2 ,......, rN , t )  0,   1, 2, 3...., k (2.6)
Block 1 The Lagrangian Formulation of Mechanics
the constraint is said to be holonomic constraint.

If not, the constraint is said to be nonholonomic. For example, a


constraint expressed as an equation containing the position as well as the
velocity coordinates, of the kind:
     
f (r1, r2 ,......, rn , r1, r2 ,......, rN , t )  0 (2.7)

Or a constraint is expressed as an inequality:


  
f (r1, r2 ,......, rN , t )  0 (2.8)

Is usually classified as nonholonomic.

So the holonomic constraint:

 is expressed as an equality

 may contain time explicitly

 does not depend on the velocity of the particles.

Typically, holonomic constraints are used to reduce the number of dynamical


degrees of freedom in a system, as you will see when you work out problems.

Going back to the example of the simple pendulum, the constraint equations
Eq. (2.2b) clearly defines a holonomic constraint since it can be written as an
equality: x 2  y 2  L0 2  0 Similarly Eq. (2.5) for the rigid body is also an
example of a holonomic constraint .

On the other hand the constraint of SAQ 1(a) cannot be written as equality
connecting the coordinates and/or time, so it is a nonholonomic constraint.

Example 2.2

Write down the equations of constraint for the following systems and classify
them as holonomic/non-holonomic:

a) a dumbbell moving in space

b) the motion of a billiard ball of radius a on a billiard table , whose length and
breadth are l and b respectively.

c) a circular disc of radius a rolling without slipping on a flat plane.

( x1, y1, z1) Solution : a) Let the coordinates of the two masses be ( x1, y1, z1) and
m1 ( x2, y 2, z2 ) and the distance between them be d (Fig. 2.5). The constraint
d on the motion of the two masses can be written as

m2 d 2  ( x 22  x12 )  ( y 2  y 1 ) 2  ( z 2  z1 ) 2 (i)
( x 2 , y 2 , z2 )
This is a holonomic constraint since it is expressed as an equality and
Fig. 2.5: Dumb bell.
depends only on the coordinates.
b) Let the coordinate of the ball be ( x, y , z ) and the origin of the coordinate
system be at one corner of the table. Since the ball has to move on the
table, the equations of constraint are:
46
Unit 2 Constrained Motion and the D’Alembert’s Principle
z0 (ii)

a  x l a (iii)

a y ba (iv)

Eq. (ii) defines a holonomic constraint, but Eqs. (iii) and (iv) are
z
nonholonomic constraints.
c) Let the position coordinates of the centre (P) of the disc be ( x, y , z )
(Fig. 2.6). Since the disc is always vertical, the z coordinate of the centre is O
always at a fixed height which is the radius of the disc, so y
P
za (v) 
x 
v
This is a holonomic constraint. The point of contact between the disc and 
the plane has the coordinates (x,y). The angle of rotation of the disc is 
and the orientation of the disc is given by the angle  .  is the angular Fig. 2.6: Vertical disc
rolling on a flat plane.
speed of the disc about its axis. To specify the motion of the disc we need
the coordinates x, y ,  and  . The condition for rolling without sliding
implies that the point of contact of the disc with the plane must be
stationary. Therefore, we must have

v  a  ; x  v cos  ; y  v sin  (vi)

Where Eq. (vi) defines the constraints on the motion which depend on the
velocity, so these are non-holonomic constraints. From Eq. (vi) we can
also write the constraints in the following form:
dx  a cos  d ; dy  a sin  d (vii)

You may also have noticed that these constraints do not depend on time,
however that may not always be the case. This brings us to another important
classification of constraints.
ii) Scleronomic and Rheonomic Constraints
Constraints defined by conditions that do not depend explicitly on time are
called scleronomic constraints. Constraints defined by conditions that
depend explicitly on time are rheonomic constraints.
An example of an equation for a scleronomic costraint would be:
     
f (r1, r2 ,......, rn , r1, r2 ,......, rN )  0 (2.9)

A typical rheonomic costraint would be:


     
f (r1, r2 ,......, rn , r1, r2 ,......, rN , t )  0 (2.10)

A scleronomic constraint can be either holonomic or nonholonomic. A


scleronomic constraint which is also holonomic would satisfy an equation
of the kind:
  
f (r1, r2 ,......, rN )  0 (2.11)
47
Block 1 The Lagrangian Formulation of Mechanics
Let us consider an example on the classification of constraints as
holonomic/nonholonomic/scleronomic/rheonomic.

Example 2.3
O
x a) Consider a simple pendulum whose length is varying with time t as
L(t)
 L  L(t )  L0 t 2 (Fig. 2.7). Write down the equation of constraint and
y
classify the constraint.

x b) Two particles are connected by a massless rigid rod of length a. The


system is moving in a plane such that the centre of the rod in the direction
y of the rod (Fig. 2.8).

Fig. 2.7: Simple Solution : a) Here the constraint equation is x 2  y 2  L0 2 t 4 which is


Pendulum. holonomic and rheonomic.
b) The coordinates of the two masses (Fig. 2.8) are ( x1, y1, z1) and
( x2, y 2, z2 ) . Clearly, since the masses move in the xy-plane, one constraint
is that
z y z1  z2  0 (i)

V Another constraint is that the distance between the two masses is fixed
a ( x1, y1, z1) and equal to a, so
P
( x1  x 2 )2  ( y1  y 2 )2  a 2 (ii)
( x 2 , y 2 , z2 )
x  x  x 2 y1  y 2 
O The coordinates of the centre of the rod (P) are  1 , ,0  . The
 2 2 

Fig. 2.8: Example 2.3(b) velocity of the centre of the rod, V  VVˆ is along the direction of the rod,
where Vˆ is the unit vector along the direction of the rod:

1

Vˆ  ( x1  x 2 )iˆ  ( y 1  y 2 ) ˆj
a
 (iii)

The velocity is

  x  x   y  y 
V  1 2 iˆ 
  1  1 2 
2 ˆj  V ( x  x )iˆ  ( y  y ) ˆj
1 2  (iv)
 2   2  a

From Eq. (iv) we can write:

x1  x 2 V y1  y 2 V
 ( x1  x 2 ) and  ( y1  y 2 ) (v)
2 a 2 a

This gives us the following constraint on the motion:

x1  x 2 y1  y 2
 (vi)
x1  x 2 y1  y 2

The constraints defined in Eqs. (i) and (ii) are holonomic and scleronomic,
but the constraint defined in Eq. (vi) is nonholomonic and scleronomic.

48
Unit 2 Constrained Motion and the D’Alembert’s Principle
SAQ 2
Write down the equations for the constraints in the following systems and
classify them:
a) A rigid rod of length L is constrained to move inside a thin walled hollow
sphere of radius A, such that both its end touch the walls of the sphere at
all times.
b) Three point masses connected by rigid rods of length l.
c) An Atwood machine with a variable length of string.

The study of constrained motion in Newtonian mechanics is difficult for the


following two reasons:
 The coordinates specifying the motion of the particles can no longer all
vary independently of each other, because they are connected by the
equations of constraint.
 Knowing or determining the constaint forces becomes an integral part of
solving the dynamical problem using the Newtonian formulation. We may
not always know what exactly these constraint forces are and they may
not be described by a known force law. As we have explained at the
beginning of this Unit, there are more unknowns than equations available
to us, even assuming that you have the equation of constraint at your
disposal. You may offcourse ask how you have solved so many problems
of constrained motion using Newtonian mechanics. When you look back
you will see that you have used equations signifying additional conditions
like the torque equations or the conservation of energy, etc in addition to
the second law. So Newton’s second law by itself would not be sufficient to
arrive at a complete solution to the dynamical problem of constrained
motion.
For example in the problem of the pendulum, while we are interested in
knowing how the pendulum moves, yet in the Newtonian formulation we have
to calculate the constraint force, which is the tension in the string before we
get to determine the motion of the pendulum.
Most often, however, we are interested only in the description of
unconstrained motion of the system. The question now is, can we now
formulate a general framework for the description of motion where:
i) we do not need to, apriori, determine the forces of constraint in the system
as we have to in the Newtonian framework, and,
ii) we work with fewer variables, which describe the unconstrained motion of
the system
In the next section we see how the these two objectives may be achieved by
using the concept of virtual work, from which we can arrive at the
D'Alembert’s Principle.

2.3 D'ALEMBERT’S PRINCIPLE


You are familiar with the concept of the work done by force acting on a particle
as it moves along a path under the action of the force. In this section we
describe the concept of virtual work.
49
Block 1 The Lagrangian Formulation of Mechanics
2.3.1 Virtual Work
We first introduce the idea of virtual displacement. A virtual diplacement, is an
infinitesimal change in the configuration of a system resulting from any
The concept of virtual 
arbitrary infinitesimal change in the coordinates of the system denoted by r ,
displacement was first
introduced by Bernoulli to at any given instant of time. This displacement is consistent with all the
explain static equilibrium forces and constraints present in the system at that instant of time. The
in 1703. displacement is called virtual because it is purely an imaginary displacement.
It has the following properties:

i) the virtual displacement r takes place at a fixed instant of time, as

opposed to a real displacement dr which take place in a time interval dt.
Therefore the applied forces as well the forces of constraint do not
change during a virtual displacement. You can imagine the system is
frozen in its motion at an instant of time, say t, and then moved without
violating any of the constraints .
ii) it is infinitesimal.

iii) the time derivatives r do not change during a virtual displacement.
 
iv) there can be a virtual displacement ri corresponding to each variable ri
being used to describe the motion.

v) the virtual displacement obeys the constraints of motion.


Suppose now, that the system we are studying is in equilibrium. Then at  any

instant of time, the total force acting on each particle must be zero or Fi  0

where Fi is the force acting on the ith particle. Therefore we can write:
 
Fi .r i  0 (2.12)

We now define the virtual work done by force Fi for the virtual displacement

ri as the quantity:
 
W i  Fi .r i (2.13)

Note
 that the virtual work is defined in exactly the same way as the real work
( Fi .dri ) except that now we have a virtual displacement in place of a real
displacement. Given the condition expressed in Eq. (2.12), we can sum over
all i to write:
 

Wi  
Fi .ri  0 (2.14)
i i
 
Let us now write the force Fi as a sum of the applied force Fi A and the force of

constraint Fi C . Then Eq. (2.14) can be rewritten as:
   
 
Fi A .r i  Fi C .r i  0 (2.15)
i i

If we now restrict ourselves to systems where the net virtual work done
by the forces of constraint in the system is zero, that is:
 
Fi C .r i  0 (2.16)
50 i
Unit 2 Constrained Motion and the D’Alembert’s Principle
The condition for the equlibrium of the system can be rewritten as: 
  N
 Fi A.ri  0 (2.17)
i

That is to say that, for the static equilibrium of the system, the virtual work s
done by the applied forces is zero. This (Eq. 2.17) is called the Principle of
Virtual Work. It is stated as follows:
The necessary condition for the static equilibrium of a dynamical (a)
system, is that the virtual work done by all the applied forces on the
system is zero, provided that the virtual work done by all the constraint O
x
forces is zero.
  
In general the applied force Fi A  0 and neither is Fi A .r i . 
T 
r
Let us discuss in some detail about the condition defined in Eq. (2.17). In
some constrained systems that you study, the constraint force on each particle
is a normal force and hence perpendicular to the displacement. In that case y
the virtual work due to the constraint force for every individual particle would (b)
 
vanish separately( Fi C .ri  0 ). Let us see some examples:

i) For a block sliding down a frictionless inclined plane(Fig. 2.9a) the force of O
 x
constraint is the normal force N and the virtual displacement is along the
   
plane. Hence F C .s  N.s  0 .
 
ii) For a simple pendulum the force of constraint is the tension in the string T T
(Fig. 2.9b). But as you can see the virtual displacement is perpendicular to
     
T and so: F C .r  T .r  0 . T y 
y 2
In that case the virtual work due to the constraint force for every individual
 
particle would vanish separately( Fi C .r i  0 ). In some systems however the
constraint force may be along the direction of displacement and in that case 
y1
the net virtual work only may be zero. This is the case for the Atwood’s (c)
       
machine (Fig. 2.9c) where F C .r  T .y 1  T .y 2  0 because y 1  y 2 Fig. 2.9: a) Bock on an
inclined plane; b) simple
Let us now work out the condition of static equilibrium for two different pendulum; and c) Atwood’s
systems using the principle of virtual work. Remember this is only for systems machine.
where Eq. (2.16) is valid.
Example 2.4 O
x

a) A mass M attached to a string of length l suspended from the ceiling of a a
 
carriage in a train which has an acceleration a , deflects by an angle  
from the vertical. Using the principle of virtual work determine the  r
 T 
magnitude of a .

b) Determine the condition for static equilibrium in an Atwood’s machine. y



Mg
Solution : a) The motion is in the xy-plane, so let us write the virtual
displacement as (Fig. 2.10): Fig. 2.10


r  xiˆ  y ˆj (i)
51
Block 1 The Lagrangian Formulation of Mechanics

The workdone by the force of constraint, which is the tension T in the

string is T .r  0 Suppose that under this virtual displacement the angle
 increases to    .Then
x  l cos   ; y  l sin   (ii)
O
x The virtual work done by the applied forces is


W   Fi A . r  Ma iˆ. l cos    iˆ  (Mg ˆj ).  l sin    ˆj 

T  Ma l cos   M g l sin  (iii)

 m2 By the principle of virtual work (Eq. 2.17), for static equilibrium W  0


y 
T y 2  a  g tan  (iv)
m1  b) Let us assume that in the virtual displacement the mass m1 moves down
 m2 g
y1 by y1 and the mass m2 moves up by y 2 , so (Fig. 2.11),
 
  y 1  y 1 ˆj ; y 2  y 1 ˆj (v)
m1g
The applied forces are m1g ˆj on the mass m1 and m2g jˆ on the mass m 2 .
Fig. 2.11: Static
equilibrium of the So the total virtual work done by the applied forces is:
Atwood’s machine.
W  m1g ˆj . y 1 ˆj  m 2 g ˆj . ( y 1 ˆj )

 m1g  y1  m2g y1  (m1  m 2 ) g y 1 (vi)

For static equilibrium, W  0, and hence we must have


m1  m2  0  m1  m2 (vii)
 
N This is the condition for static equilibrium of the Atwood’s machine.
dr2

dr
 The condition on the virtual work done by the constraint forces (Eq. 2.16) is
dr1
true for most constrained systems you study, in one form or the other.
However you may have noticed that in these systems the real work done by
the forces of constraint is also zero. So you may wonder why we need a
Fig. 2.12: Block sliding principle of virtual work, using a virtual displacement.
down a moving inclined
plane. It is because there are some situations where the real work done by the
constraint forces is not zero. Let us look at some such systems. This is usually
O
in systems where we have time varying constraints of the kind shown in
x
Figs. 2.12 and 2.13:
i) In Fig. 2.12 you see a block sliding down a moving frictionless plane. Now
  because the inclined plane is moving, in the time interval t to t + dt, the
T dr   
 actual displacement of the particle dr  dr1  dr 2 has a component
r 
normal to the surface, that is, in the direction of the force of constraint ( N ).
y In this situation, the real work done by the force of constraint is not zero.
But, for a virtual displacement, dt =0. Therefore, the virtual displacement
Fig. 2.13: Simple
of the particle is still along the plane and the virtual work done by the
pendulum with varying
length. force of constraint will be zero.

52
Unit 2 Constrained Motion and the D’Alembert’s Principle

ii) In Fig. 2.13 you see a simple pendulum with a length that varies with time.
  
Here the real work done by the constraint force T ( T .dr ) is not zero,
  
though the virtual work done by T ( T .r ) is zero.

SAQ 3
Calculate the virtual work for a small bead of mass m sliding under gravity on a
fixed frictionless elliptical wire.

Let us now study how the principle of virtual work provides us with an extra
set of equations, to add to the 3N equations of motion derived from Newton’s
second law for a constrained system. This happens only if the virtual
displacements are independent of each other.
2.3.2 D'Alembert’s Principle
Notice that while the principle of virtual work, solves a part of our problem in
that it does not contain the forces of constraint, but it is still only a condition of
static equilibrium. We have yet to arrive at the dynamics of the system. Let us
see how that can be done.
  
The equation of motion for the ith particle, Fi  p i , where Fi is the net force

acting on the ith particle and pi is its linear momentum, can also be written as :
 
Fi  p i  0 (2.18)

Eq. (2.18) can also be stated as follows:


Each particle of the system is in equlibrium under the action of the
 
following two forces acting on it: Fi and  p i .
  
While the force acting on each particle Fi  Fi A  Fi C , is the sum of the

applied and constraint forces,  p may be thought of an effective reverse
i
force, sometimes called the inertial force, which would bring the system to
equilibrium. From Eq. (2.18) we can also write:

 
  
Fi  p i .r i  0 (2.19)
i

We already know that the virtual work done by force Fi for the virtual
      
displacement r i is: W i  Fi .r i  Fi A .r i  Fi C .r i

As we have already studied in the previous section, the total virtual work is
     
W  
W i  
Fi A .r i  
Fi C .r i  
Fi A .r i (2.20)
i i i i
 
since  Fi C .r i   0 . Eq. (2.19) therefore reduces to:
i

 Fi A  p i . ri
  
0 (2.21)
i

This is called the D'Alembert’s principle. The physical significance of this


principle is that it tells us that the momentum of the particle is determined only
53
Block 1 The Lagrangian Formulation of Mechanics
   
by the Fi A and not by Fi C . Remember that the statement is not Fi A  p i  0 ,
which, in general, is not true for a constrained system. It is true only when
multipled by the virtual displacement. Notice also, that Eq. (2.21) is different
from the Principle of Virtual Work, because it includes the dynamics of the

system through the term p i . In particular,

i) The forces of constraint do not figure in the equation in any way, all you
need to know are the applied forces.

ii) The constraint relation enters the equation only through the term r i ,
through the relation between the coordinates.
iii) The only condition required to be satisfied by a dynamical system for the
D'Alembert’s principle to be applicable is that the virtual work done by all
the constraint forces vanishes (Eq. 2.16).
The D'Alembert’s principle is a formulation for determining the dynamics of the
system which does not require a knowledge of the constraints. So one has
solved a part of the problem. Yet Eq. (2.21) cannot really be applied to derive
the equatins of motion of a constrained system, because the virtual

displacements ri are not all independent. You know that in a constrained
system, there are equations of constraint which connect the some of the
 
variables ri . If r i were independent variables, we could have written:

 
 
Fi A  p i . r i  0 for i  1, N (2.22)

This would then be a complete description of the constrained dynamical


system, where we have 3N differential equations for 3N unknowns (unlike
Newton’s second law). However, to actually use Eq. (2.21) to determine the
motion of a system, one has to eliminate those degrees of freedom in the
system which have no dynamics associated with them. We do this using the
constraints in the system.

SAQ 4
Show that the D'Alembert’s principle is also a statement of the conservation of
energy, if the virtual displacements are replaced by real displacements.

In the next section we introduce the concept of generalized coordinates as the


next step towards arriving at a general formalism for the study of the motion of
constrained systems.

2.4 GENERALIZED COORDINATES


We have defined the degrees of freedom of a dynamical system in Sec. 1.1.
To repeat, it is the number of independent quantities required to completely
specify, in a unique way, the position of the particles of the system. In the
absence of any constraints, an N-particle system has 3N degrees of freedom
and the number of independent coordinates required to specify the
configuration of the system is 3N.
Note that the dynamical degrees of freedom of a system do not depend on the
choice of coordinates being used to describe the system. The only restriction
54 is that the number of independent coordinates should be equal to the number
Unit 2 Constrained Motion and the D’Alembert’s Principle
of degrees of freedom. So far, we have not really discussed what these
independent coordinates should be, that describe the configuration (position of
each particle) of the system uniquely. However just knowing the configuration
of the system at any instant of time, does not tell us its configuration at all
times. For that we need to solve the equation of motion derived from the
second law of motion, central to the Newtonian formulation of mechanics:
 
mr  F
There is no restriction in the choice of independent coordinates that can be
used to describe the motion of the system, but this differential equation is most
conveniently expressed in rectangular Cartesian coordinates. For example,
for single particle motion, using a rectangular Cartesian coordinate system,
this equation reduces to the following three equations for the coordinates
x(t ), y (t ), z(t ) :
mx(t )  Fx ; my(t )  Fy ; mz(t )  Fz (2.23)

In general, these are not independent equations and may be coupled through
the force depending on all three position coordinates and/or the components
of the velocity. Recollect, further, that in your earlier courses in mechanics
you have often solved problems in mechanics using non-Cartesian
coordinates because it was more convenient to do so given the nature of
force, etc: for example in central force motion.
The general approach while using non-Cartesian coordinates has always been
the following:
i) Write the transformation equations connecting the Cartesian (x,y,z)
coordinates and the non-Cartesian coordinates say (u1,u2,u3):
x (t )  x (u1(t ), u 2 (t ), u 3 (t )) (2.24a)
y (t )  y (u1(t ), u 2 (t ), u 3 (t )) (2.24b)
z(t )  z(u1(t ), u 2 (t ), u 3 (t )) (2.24c)
For example, to describe the motion of a particle in a plane, the position
can be specified either in rectangular Cartesian coordinates (x,y) or by
in plane polar coordinates(r,). In three dimensions the we could use
either rectangular Cartesian coordinates (x,y,z), or cylindrical
coordinates (ρ,ϕ,z) or spherical polar coordinates (r, ,ϕ). The
transformation equations connecting the Cartesian coordinates to the
spherical polar coordinates are:
x (t )  r sin  cos  (2.25a)
y (t )  r sin  sin  (2.25b)
z(t )  r cos  (2.25c)
ii) Obtain the equations of motion in the new coordinates. For that we first
determine the differential elements of the Cartesian coordinates
(dx, dy , dz ) in terms of the differential elements of the new coordinates
(du1, du 2 , du3 ) and then obtain  x(t ), y(t ), z(t ) in terms of
u1(t ),u2 (t ),u3 (t ).
 
iii) Next we rewrite mr  F in terms of the new cordinates using the
transformation equations and the expressions for  x(t ), y(t ), z(t ) in terms
of u1(t ),u2 (t ),u3 (t ) .
55
Block 1 The Lagrangian Formulation of Mechanics
Unlike Cartesian coordinates, these new coordinates do not all have the
dimension of length, although they can be used to specify position in a
unique way.
At this stage we introduce the term generalized coordinates. If a dynamical
system has n degrees of freedom, then any set of n independent coordinates
that completely define the position of each particle in the system are called its
generalized coordinates. The term generalized is used to emphasize that
these coordinates need not be the usual Cartesian coordinates which are
the position coordinates having units of length. The generalized coordinates
are not necessarily lengths or angles, but may be any quantity appropriate to
the description of the position of the system. The only restrictions on the
generalized coordinates are
 The generalized coordinates are independent of each other
 The number of generalized coordinates is equal in number to the degrees
of freedom of the system.
For an unconstrained N particle system, we could define a set of 3N
generalized coordinates q1, q2 ,.......q3N connected to the 3N Cartesian
coordinates.
In defining the At this point, for the sake of simplicity in notation, we write the Cartesian
Cartesian coordinates
we have shifted from
coordinates of the N-particle system as:
 x, y , z  to an  x1(t ), x2 (t ), x3 (t ),  x 4 (t ), x5 (t ), x 6 (t ),.....,  x3N  2 (t ), x3N 1(t ), x3N (t )
 x1, x2 , x3  notation in place of the familiar:
for simplicity. So the x1(t ), y 1(t ), z1(t ), x 2 (t ), y 2 (t ), z2 (t ),....., x N (t ), y N (t ), zN (t ) .
Cartesian coordinates
of particle 1 are So now we have 3N Cartersian coordinates for the N-particle system
 x1, x2 , x3  , that of { x i }, i  1,2....3N (read the margin remark) which are related to the 3N
particle 2 are generalized coordinates q1, q2 ,.......q3N as:
 x4 , x5 , x6  and so on.
x1  x1q1, q2 ,.......q3N , t 

x 2  x 2 q1, q 2 ,.......q3N , t 

x3  x3 q1, q 2 ,.......q3N , t  (2.26)


x3N  x3N q1, q2 ,.......q3N , t 

Notice that the number of generalized coordinated is equal to the number of


rectangular Cartesian coordinates in Eq. (2.26).
You have seen how the number of degrees of freedom reduces with the
presence of constraints in the dynamical system. This also reduces the
number of coordinates needed to completely specify the system. Let us see
how.
The condition for a set of 3N variables (say x1, x2,..., x3N ) to be independent
is the following:
3N
 c i dx i  0  c1  c 2  ...  c 3N  0 (2.27)
i 1

56
Unit 2 Constrained Motion and the D’Alembert’s Principle
In other words, it is not possible to write any one differential element as a
linear combination of the other differential elements. This is possible only
when there is no (constraint) relationship connecting these variables.
For example, for the simple pendulum, the constraint equation is
x 2  y 2  L2  dx  dy  0 or dx  dy (2.28)

So x and y are no longer independent variables, and the constraint equation


can be used to eliminate either one of the variables (x or y) so that we are left
with just one independent variable. The system therefore does not have two
degrees of freedom, but just one. The constraint therefore has the effect of
reducing the number of degrees of freedom in the system.
In general, suppose that in an N particle system, there are a number k of
constraints that can be expressed as equations of the form given by Eq.(2.2),
i.e, we consider a dynamical system which has k holonomic constraints.
We can use the constraint equations to eliminate k out of the 3N coordinates
we started with, so that we are left with just 3N  k coordinates, which are
now independent. Effectively, the system will then have 3N  k degrees of
freedom.
The next question is: how does one choose these 3N  k independent
coordinates? One obvious way is to pick a set of 3N  k coordinates from the
3N Cartesian coordinates we started with. In this case the 3N  k Cartesian
coordinates would be the generalized coordinates for the system ( fewer in
number than the original 3N Cartesian coordinates). However this choice of
coordinates is not binding. We can also choose any other set of 3N  k
coordinates q1, q 2 ,.......q 3N k , and each of the original 3N Cartesian
coordinates may be expressed in terms of these coordinates as follows:
x1  x1q1, q2 ,.......q3N  k , t 

x 2  x 2 q1, q 2 ,.......q3N  k , t 

 (2.29)
x3N  x3N q1, q2 ,.......q3N  k , t 

For the constrained system therefore, the minimum number of


independent coordinates (Cartesian or otherwise) needed to define
completely the configuration of the entire system, while taking
advantage of the constraints on the system are the generalized
coordinates. The generalized coordinates need not be the rectangular
Cartesian coordinates of the system.
In Sec 2.2.1 we have already explained that for a simple pendulum with a
string of length L0 , the system has two constraints z  0 and x 2  y 2  L0 2 ,
both of which as you can see are holonomic constraints. Therefore the simple
pendulum has just one (number of position coordinates of the corresponding
unconstrained system  number of holonomic constraints =3-2=1) dynamical
degree of freedom, and therefore we need just one generalized coordinate to
describe its motion. We could select either of the Cartesian coordinates x or y
as the generlized coordinate (not z, since z  0 ). In that case the
transformation equations (Eq. 2.29) would be: 57
Block 1 The Lagrangian Formulation of Mechanics
2
x  x; y  L0  x 2 ; z  0 (2.30a)

The choice of generalized coordinates for a dynamical system is not unique.


We could also choose the angle  (Fig. 2.1) as the generalized coordinate, in
which case the transformation equations are:
x  L0 sin ; y  L0 cos ; z  0 (2.30b)

Now note that:


A O B i) The number of generalized coordinates for a constrained system is less
x than the number of independent coordinates used to define the position
a
vectors for each particle in the corresponding unconstrained system. Also
y2 in an N particle system, it will not be possible, in general to club the
y1 generalized coordinates in sets of three, with one such set for each
different particle.
m2
y ii) To solve for the dynamical variables of the system of particles, we now
set up equations of motion in the generalized coordinates and solve them.
m1
Then we write down the position vectors for each particle using the
(a) transformation equations of Eq. (2.29).

O Let us now try to write down the generalized coordinates for some of the
x
A mechanical systems you have studied in this chapter so far.
LA
yA Example 2.5
yB
xA Identify generalized coordinates for the following systems:
A
B LB a) An Atwood’s machine (Fig 2.14a) where the radius of the pulley is a and
the length of the string connecting the two masses is l.
B
xB b) A double pendulum as defined in Example 2.1 (Fig. 2.14b).
c) A block sliding down an inclined plane (Fig. 2.14c).
y
(b) Solution : a) There are 5 holonomic constraints:
y
x1  x 2  z1  z2  0

s and y1  y 2  a  l
 where  a is length of the string between the points A and B. So the system
a
has one degree of freedom (3 2  5=1) and hence we need just one

generalized coordinate which we can chose to be y1 or y 2 with the
transformation equations being y 2  l  a  y1 and y1  l  a  y 2.
b x
b) The four holonomic constraints defined in Eqs. (i), (ii) and (iii) of
z
(c) Example 2.1 indicate that there are just two dynamical degrees of freedom
(3 2  4=2) and hence there should be just two generalized coordinates.
Fig. 2.14: a) Atwood’s Let us choose these coordinates to be,  A and  B . Thus
machine; b) double
pendulum; c) block on x A  L A sin  A ; y A  L A cos  A ; z A  0
an inclined plane. and
xB  LA sin  A  LB sin B ;

y B  LA cos  A  LB cos B ; zB  0

58
Unit 2 Constrained Motion and the D’Alembert’s Principle
c) The block (coordinates x,y,z) follows a straight line path down the incline in
the xy-plane( z  0 ).
y
And also : tan  
bx
ay
Alternatively, we could write tan   . In either case, the system has
x
two holonomic constraints and hence the block has just one degree of
freedom and one generalized coordinate, which we can choose as s
(Fig. 2.14c), the distance of the block from the top of the incline. We get:
x  s cos  ; y  a  s sin  ; z  0

Remember that although the choice of generalized coordinates is not unique,


yet a certain set of generalized coordinates may sometimes make the
solution simpler.
You may like to write down the generalized coordinates for a few systems on
your own.

SAQ 5
Identify generalized coordinates for the following dynamical systems:
a) A simple pendulum of mass m suspended from a support of mass M which
is moving to the right .
b) A block of mass m sliding down a frictionless incline which in turn is sliding
on a smooth horizontal plane which has a mass M.
c) A spherical pendulum as defined in SAQ 1d.
A O B
x
Note that the entire discussion on generalized coordinates and effective
degrees of freedom is related to holonomic constraints only. y2

Once we introduce the concept of generalized coordinates, D'Alembert’s y1 T
principle can provide us with a complete solution of the dynamics of a
 m2
mechanical system. Let us see how, by solving two examples using 
T y
D'Alembert’s principle. y 2

Example 2.6 m1 m2 g ˆj

y1
Using D'Alembert’s Principle:

a) The Atwood’s machine (Example 2.4b). m1g ˆj


Fig: 2.15
b) Derive the equations of motion for the block and the plane for the system
of SAQ 5b

Solution : a) The system has just one degree of freedom. Let the
generalized coordinate be y1 which is the downward displacement of the

mass m1 (Fig. 2.15). The virtual displacement of the two blocks are y 1
and y 2 . Note that while the mass m1 moves down by y1 the mass

59
Block 1 The Lagrangian Formulation of Mechanics
 
m2 move up by y1 therefore y 1  y 1 jˆ  y 2 The applied forces are
 
F A  m1g ˆj acting an m1 and F A  m 2 g ˆj acting an m 2 .
1 2

 D'Alembert’s principle gives us:


     
(F1  p 1 ). y 1  (F2  p 2 ). y 2  0 (i)
 
where p 1  m1y1 ˆj and p 2  m 2 y1 ˆj (ii)
So we get: (m1g  m 2 g  m1y1  m 2 y1 ) y 1  0
which gives us:
 m  m2 
y1   1 g
 m1  m 2 

b) We have already described the generalized coordinates for this system in


SAQ 5(b). Since there are two generalized coordinates,

s and X, there are

two independent virtual displacements s and X (Fig. 2.16). The virtual
 
work done by the forces of constraint is N1.s  N 2 .X  0.

y

s N1
 
A s a

Fblock s
X   
X X
N2 
x
A
Fincline

Fig. 2.16: Block sliding down a moving incline.



The displacement s is an increase in s by s , at a fixed time and
involves only the motion of the block down the incline. The virtual

displacement of the block is s . The applied force acting on the block is

the gravitational force acting on the block, F A  mg ˆj and the inertial
block
  
force acting on the block is p block  ma where a is the acceleration of

the block. a is the vector sum of the acceleration of the block down the

incline, s , and the acceleration of the block along the flat plane on which
      
the incline is moving, X (Fig. 2.16). So a  s  X and p block  m s  X .  
From D'Alembert’s principle we get:
     
(F A  p block ). s    mg ˆj  m s  X  . s  0
block   
We get the following equation for the acceleration of the block down the
incline:
 cos   s  g sin   X
mg sin   ms  mX  cos 

Notice that for a stationary inclined plane  X


  0  , this reduces to

s  g sin 

We next consider the virtual displacement X in which X increases by X ,
60 and there is a displacement of the system comprising the block and the
Unit 2 Constrained Motion and the D’Alembert’s Principle
incline by X . The applied force is the gravitational force,
A
ˆ
Fblock  incline  (M  m )g j and the inertial force is
  
p block  incline  ms  (m  M ) X . D'Alembert’s Principle then gives us:
 
 mg ˆj  Mg ˆj  m s  X  
  MX  . X  0
   

So the the acceleration X  is

MX   m X
  m s cos   0  X    m s cos 
(M  m )

In the next Unit we will study how using the concept of generalized
coordinates in the D Álembert’s principle, leads us to the Lagrangian
formulation of mechanics.

2.5 SUMMARY
 Constraints

When the motion of a dynamical system is constrained in some way,


the particle or system of particles is restricted to move in a particular
way, such that the coordinates of the particle satisfy certain given
conditions.

The prescribed path or surface along which motion takes place is


called a constraint The equations that define the path or surface
along which the motion takes place are called the constraint
equations.

The forces that are responsible for creating these constraints on the
motion of the particle/particles in the system are called the forces of
constraint.

Constraints reduce the number of dynamical degrees of freedom of a


system.

 Classification of Constraints

i) Holonomic and Nonholonomic Constraints

For an N-particle system described by the set of position vectors


  
r1, r 2 ,......, rN , if the constraint is defined by an equation connecting
the coordinates of the particles (and possibly time), of the following
form:
  
f  (r1, r 2 ,......, r N , t )  0,   1, 2, 3....,3N  k (2.6)

the constraint is said to be holonomic constraint. Holonomic


constraints are independent of the velocities. If the constraint
relation is not holonomic, it is said to be nonholonomic. A 61
Block 1 The Lagrangian Formulation of Mechanics
nonholomic constraint may depend on the velocities of the particles
in the system or be expressed as an inequality.

ii) Scleronomic and Rheonomic Constraints

Constraints defined by conditions that do not depend explicitly on


time are called scleronomic constraints. Constraints defined by
conditions that depend explicitly on time are rheonomic constraints.

 Virtual Displacement

A virtual diplacement, denoted by r , is an infinitesimal change in the
configuration of a system resulting from any arbitrary infinitesimal
change in the coordinates of the system, at any given instant of time.
This displacement is consistent with all the forces and constraints
present in the system at that instant of time.

 Virtual Work

If Fi is the net force acting on the ith particle of the system, we define
 
the virtual work done by force Fi for the virtual displacement r i as
the quantity:
 
W i  Fi .r i
 
The total virtual work in the system is  W i   Fi .r i  0.
i i

In a constrained system the net force is the sum of the applied force
and the force of constraint.
The Principle of Virtual Work states that:
The necessary condition for the static equilibrium of a
dynamical system, is that the virtual work done by all the
applied forces on the system is zero, provided that the virtual
work done by all the constraint forces is zero:

 
 Fi A .r i 0 (2.17)
i

where Fi A is the applied force on the ith particle of the system.

 D'Alembert’s Principle
D'Alembert’s Principle states that for an N-particle system where the
virtual work done by all the constraint forces is zero we can write:

 Fi A  p i . ri
N   
0
i 1

where  p i may be thought of as an effective reverse force, also called
the inertial force.
 Generalized Coordinates
62
Unit 2 Constrained Motion and the D’Alembert’s Principle
If a dynamical system has n degrees of freedom, then any set of n
independent coordinates that completely define the position of each
particle in the system are called its generalized coordinates.
An N particle system with k holonomic constraints has 3N  k
degrees of freedom and hence 3N  k generalized coordinates.  

2.6 TERMINAL QUESTIONS Fig. 2.17: TQ 2.


y
1. Calculate the virtual work for a small block of mass m moving down an
frictionless inclined plane of mass M moving to the right on a smooth m
surface(Example 2.6). The inclined plane makes an angle  with the
horizontal. 
 m F
x
2. Two unequal masses m1 and m2 are constrained to move on two smooth
inclined planes and they are connected by an inextensible string of length Fig. 2.18: TQ 3.
L which passes over a fixed smooth pulley P (Fig. 2.17). Determine the
condition of static equilibrium.
3. Two frictionless blocks of mass m each are connected by a massless rigid
rod of length L. The system is constrained
 to move in the vertical plane
under the action of an applied force F as shown in Fig. 2.18. Calculate the
condition for static equilibrium.
4. Use D'Alembert’s Principle to write the equations of motion (i) a body of
mass m falling freely under gravity, (ii) a simple pendulum and (iii) a box of y
mass m on a stationary inclined plane.
5. Use D'Alembert’s Principle to write the equations of motion for the system
described in TQ2.
( x, y , z )

6. Use D'Alembert’s Principle to write the equation of motion moving for a r
bead of mass m moving on a frictionless vertical hoop of radius R under
x
the action of gravity.
R
7. Use D'Alembert’s Principle to write the equation of motion for a spherical z
pendulum of mass m with a variable length given by r  R(1  cos t ) .
Fig. 2.19: SAQ 1a.
2.7 SOLUTIONS AND ANSWERS
z
Self-Assessment Questions ( x1, y1, z1)
O x
1. a) The particle moves on the surface of the sphere till it falls off the L0
sphere. Taking the coordinates of the particle to be (x, y, z) where the
origin of the coordinate system is the centre of the sphere(Fig. 2.19),
the equation of constraint is: ( x2, y 2 , z2 )
 y
r  R or x 2  y 2  z 2  R 2 Fig. 2.20: Simple
pendulum with a
b) Let the x-axis lie along the direction of motion of the support moving support.
Fig. (2.20). The coordinates of moving support are ( x1, y1, z1) and that
of the bob of the pendulum are x2 , y 2 , z2  .The equations of constraint
are z1  0; y1  0; z2  0 and ( x 2  x1)2  y 22  L0 2 .

63
Block 1 The Lagrangian Formulation of Mechanics
y c) The coordinate of the bead are x, y , z  and the wire is in the xy-plane
(Fig. 2.21). The equations of constraint are:
(x, y, z) x2 y2
b z  0 and   1.
x a2 b2
a
d) The coordinates of the bob are x, y , z  (Fig. 2.22). However, the motion
z pendulum as confined to the surface of a sphere of radius L0 , so the
Fig. 2.21: SAQ 1c. equation of constraint is:

y x 2  y 2  z 2  L0 2
( x, y , z )
2. a) Let us coordinates of the two ends of the rod be  x1, y 1, z1  and
x2 , y 2 , z2  (Fig. 2.23). Since both ends of the rod must always touch
x
the inner wall of the sphere, we must have:
L
z x12  y 12  z12  A 2 ; x 22  y 22  z 22  A 2 (i)
Fig. 2.22: SAQ 1d.
And since the length of the rod is L, we must have
y
( x 2  x1 ) 2  ( y 2  y 1 ) 2  ( z 2  z1 ) 2  L2 (ii)
( x 2 , y 2 , z2 )
All three constraints are holonomic and scleronomic.
A x b) The constraints are (Fig. 2.24):
L
z ( x1  x 2 ) 2  ( y 1  y 2 )2  ( z1  z 2 ) 2  l 2
( x1, y1, z1)
( x 2  x 3 ) 2  ( y 2  y 3 ) 2  ( z 2  z3 ) 2  l 2
Fig. 2.23: SAQ 2a.
( x 3  x1 ) 2  ( y 3  y 1 )2  ( z3  z1 ) 2  l 2
y
The constraints are holonomic and scleronomic.
( x3 , y 3 , z3 )
c) For an Atwood’s machine (Fig. 2.2) with variable length, say l  l (t ) , in
l addition to the constraints z1  z2  x1  x 2  0 which are holonomic
x
( x 2 , y 2 , z2 ) and scleronomic, we also have the constraint y 1  y 2  l (t ) which is
z ( x1, y1, z1) holonomic and rheonomic, since the length depends explicitly on time.

Fig. 2.24: SAQ 2b. 3. The virtual displacement is r  x iˆ  y ˆj , (the wire is in the xy-plane,
y hence z  0 ) tangential to wire (Fig. 2.25). So the forces are the

 constraint F C which is force exerted by the wire to keep the bead on the
( x, y , z ) F C 
 wire, which is perpendicular to the virtual displacement r and the
 r  
r applied force, which is the weight of the bead : F A  mg ˆj . The virtual
b FA
x
a work done is
   
W  F A . r  F C . r   mg y
z
Fig. 2.25: SAQ 3. 4. From D'Alembert’s principle we have

 
  
Fi A  p i . ri  0
i
The virtual displacement can be any arbitrary displacement that is
consistent with the constraints in the system. Let us say that it is a real
infinitesimal displacement. So,

64
Unit 2 Constrained Motion and the D’Alembert’s Principle

 
  
Fi A  m i ri . dri  0 (i)
i
We can simplify the second term as follows:
  d 1   dT i
 mi ri .dri  dt  2 mi ri .ri  dt   dt
dt  dT
i i i

where T is the kinetic energy of the system. Let us now assume that the
applied force in the system is conservative and can be written as the
   
gradient of a scalar quantity and so Fi A . dri  
  i V . dri  dV .
i i
Eq. (i) is then
dV  dT  d (T  V )  0  T  V  Constant
5. a) As explained in SAQ 1(b) the coordinates of the support are ( x1, y 1, z1 ) z
( x1, y1, z1)
and that of the bob of the pendulum are ( x 2 , y 2 , z 2 ) and the motion of O x
the support is along the x-axis.The system has two degrees of freedom L0

since there are four constraints and so we need two generalized
coordinates to describe the motion of the system. We can choose
these to be the x-coordinate of the moving support, x1, and the angle , ( x2, y 2, z2 )
y
shown in Fig. 2.24. The transformation equations (Eq. 2.29) are:
Fig. 2.26: SAQ 5(a).
x1  x1; y 1  0; z1  0
and
x2  x1  L0 sin ; y 2  L0 cos ; z2  0
b)
y

s

a

X

x
b
z
Fig. 2.27: Block sliding down a moving incline.
We consider the block and inclined plane system of Example 2.5(c), in
which the plane is now moving to the right along the x-axis. The
generalized coordinate for the motion of the block on the incline is s,
which is the distance of the block from the top of the incline. Since
there is no movement of the plane in the y or z direction, we need just
one generalized coordinate to describe the motion of the plane along
the x-axis, say X as shown in Fig. 2.27.
c) The system has two degrees of freedom so can choose the
generalized coordinates to be the polar angle  and azimuthal angle  .

Terminal Questions
1. We calculate the virtual work done for the motion of the block down the

inclined plane, for the virtual displacement s . As explained in Example

2.6, the displacement s is an increase in s by s , at a fixed time
(Fig.2.16) . This involves only the motion of the block down the incline. The
65
Block 1 The Lagrangian Formulation of Mechanics
 
virtual displacement of the block is rblock  s . The applied force acting

on the block is the gravitational force acting on the block, F A  mg ˆj ,
block
 
the force of constraint is the normal force F C  N1 which is
block

perpendicular to s . The virtual work done is:
    
W  (F A  F C ). s  ( mg jˆ  N1 ). s  mg sin  s
block block

2. We use the principle of virtual work to solve this problem. The virtual
 
displacements are s1 for mass m1 and s 2 for mass m2 (Fig. 2.28).The
constraint on the motion is s1  s 2  L, so the system has just one degree
of freedom and we choose the generalized coordinate to be either s1 or
s2 . y P

 
N1 N2

s1 s s2 
 1 
F1 s2
F2
 
x
Fig. 2.28: TQ 2.

The forces are the gravitational force F1  m1g jˆ and the normal reaction
 
force N1 acting on the mass m1 and the gravitational force F2  m2 g ˆj

and the normal reaction force N 2 are acting on the mass m2 .
   
Work done by forces of constraint: N1 .s1  N 2 . s 2  0 .
Therefore, using the principle of virtual work we can write:
  
F1. s1  F2. s2  0  m1g sin  s1  m2 g sin  s 2  0 (i)
Since s1  s 2  L  s1  s 2 , Eq. (i) reduces to

m1g sin   m2 g sin 


which is the required condition for static equilibrium.
3. y


 N1
r1  
F1 N2 
r2 
 F
 x
F2

Fig. 2.29: TQ 4.
 
The virtual displacements are r1  yˆj and r2  xiˆ as shown in
Fig. 2.29.
 The work done by the constraint forces is
  
zero (N1 .r1  N 2 . r2  0) , so using the principle of virtual work we get :
     
F1. r1  F2 . r2  F . r2  0
 
where F1  mg jˆ , F2  mg ˆj . So,
mg y  0  F x  0  mg  y  F  x z
66


Unit 2 Constrained Motion and the D’Alembert’s Principle

Since the length of the rod is fixed, we have x 2  y 2  L2 , which gives us: z

xx  yy  0  xx   yy



y y z F
Also:  tan  so   cot  1
x x
y
So we get F  mg cot 
x
4. i) The motion is one-dimensional and the  generalized coordinate is Fig. 2.30: TQ 4(i).
z (Fig. 2.30). The applied is force  F1  mg kˆ and the inertial force is

p  mz kˆ .

The virtual displacement is z  zkˆ .Using D'Alembert’s principle we
write:
O
 
   x
F1  p .z  0  mg  mz z  0 L0

Which give us the familiar equation of motion z  g. 

ii) The generalized coordinate is , the virtual displacement in the r
 F1
direction of motion is r  L0ˆ where ̂ is the unit vector in the
 y
direction of increasing  (Fig. 2.31). The applied force is F1  mg ˆj and
 Fig. 2.31: TQ 4(ii).
the inertial force is p  m L  . From D'Alembert’s principle we get:
0
   
F1. r  p.r  0  ( mg L0 sin   mL0 2)   0

g
The equation of motion is:    sin .
L0

iii) For a block on a fixed inclined plane (Example 2.5c) the generalized
coordinate is s, the applied force is F1  mg ˆj , where m is the mass of
 
the block and the inertial force is p  ms . The virtual displacement is

s . With D'Alembert’s principle we get
   
F1.s  p.s  0  mg sin  s  mss  0

The equation of motion is s  g sin  .

5. Continuing from the solution of TQ 2 (Fig. 2.28), we can write the inertial
   
forces on the two masses m1 and m2 as p1  m1s1 and p 2  m2 s2
respectively. From D'Alembert’s Principle we have:
       
F1.s1  F2 . s2  p1.s1  p 2 .s2  0
 
 m1 g sin  s1  m2g sin  s2  m1s1 s1  m2s2  s2  0

However, since s1  s 2  L , s2  s1 and s1  s2 . Using s1 as the


generalized coordinate, the equation of motion for s1 is

m g sin   m 2 g sin 
s1  1
m1  m2
6. Let the coordinates of the bead be ( x, y , z ) . The constraints on the motion
of the bead are:
67
Block 1 The Lagrangian Formulation of Mechanics
i) z  0, since the loop is fixed vertically in the xy-plane.

ii) x 2  y 2  R 2

Thus there is just one generalized coordinate, which we can take as the
z angle  shown in Fig. 2.32.The applied force is F1  Mg iˆ , the inertial
 
force is p  m R ˆ and the virtual displacement is s  R   ˆ , where ̂ is
the unit vector in the direction of increasing  .

s Therefore, using D'Alembert’s principle we get,
 F1    
x F1.s  p.  s  0  (mg R sin   m R 2)    0
R g
And the equation of motion is:   sin .
R
z
7. As discussed in SAQ 5c, the generalized coordinates for a spherical
Fig. 2.32: TQ 6. pendulum of fixed length are  and  . In this question the length is
variable, so we have three generalised coordinates r , , , which are the
spherical polar coordinates. In spherical polar coordinates the acceleration
of the system is

r  ( r  r 2  r 2 sin 2 )rˆ  (r  2r  r  2 sin  cos ) ˆ

 ( r sin   2r  sin   2r  cos ) ˆ



The virtual displacement is: r  r  ˆ  r sin   ˆ ( for a virtual
displacement , dt is zero and hence r  0 ).

The applied force is F  mg cos rˆ  mg sin  ˆ

From D'Alembert’s principle


   
F .  r  m r. r  0

 ( mg cos  rˆ  mg sin  ˆ ).( r   ˆ  r sin    ˆ )

 m (r  r  2  r 2 sin 2 ) rˆ . (r   ˆ  r sin   ) ˆ

 m (r  2r   r  2 sin  cos )ˆ .( r ˆ  r sin  ˆ )

 m (r  sin   2r  sin   2r   cos ) ˆ . (r   ˆ  r sin    ˆ )  0

On simplifying these equations we get the following equations of motion for


 and  :

r   2r   r  2 sin  cos   g sin 

and r sin    2r  sin   2r cos     0

The equation of motion for r is: r   R  sin t  r  R2 cos t

68
Unit 3 Lagrange’s Equations and its Applications

UNIT 3
LAGRANGE’S
EQUATIONS AND ITS
APPLICATIONS
Structure

3.1 Introduction 3.4 Euler-Lagrange Equations of


Expected Learning Outcomes Motion for Simple Systems
3.2 Lagrange’s Equations from Unconstrained Motion of a Single
D'Alembert’s Principle Particle
Lagrange’s Equations of the Second Kind 3.5 Summary
Lagrangian 3.6 Terminal Questions
Kinetic Energy in Generalized Coordinates 3.7 Solutions and Answers
3.3 Lagrange’s Equations for a Velocity-
dependent Potential

3.1 INTRODUCTION
In Unit 2 you have seen that in the presence of constraints, the equations of
motion derived from Newton’s second law are no longer sufficient to obtain a
solution for the dynamics of a system. You also studied, how, using the
D'Alembert’s principle and generalized coordinates, one can derive the
equations of motion for some simple systems without having to determine the
forces of constraint as is essential in the Newtonian formulation.
D'Alembert’s principle is, in fact, one of the most fundamental principles in
physics. As you have seen in Unit 2, it is also a statement of conservation of
energy and the cornerstone of classical, relativistic, and quantum mechanics. In
the words of Cornelius Lanczos “All the different principles of mechanics are
merely mathematically different formulations of D 'Alembert’s Principle”
(The Variational Principles of Mechanics, Dover Publications, INC., New York,
1949). As you study more of classical, relativistic and quantum mechanics, you
will understand the scope of this simple principle. In this unit we derive the
Lagrangian formulation of mechanics starting from the D'Alembert’s principle.
The Lagrangian formulation of mechanics is a much cleaner way of solving
problems in mechanics, especially those of constrained motion. It also holds for a
surprisingly large class of constrained mechanical systems. Conceived by Italian
born, French mathematician Joseph Louis Lagrange (1736 – 1813) at the age of
23, it is an analytical method that allows us to arrive at a sufficient number of
independent equations to help us solve a dynamical problem. The Lagrangian 69
Block 1 The Lagrangian Formulation of Mechanics
formalism introduces an important scalar quantity the “Lagrangian” of a
mechanical system. Unlike the Newtonian formulation in which we work with the
forces acting on a system, which is a vector quantity, in the Lagrangian
formulation the equations of motion of the system are derived from a purely
scalar quantity, the Lagrangian. The Lagrangian also is the central quantity in
another fundamental principle of physics that you will study in Unit 4, namely the
Hamilton’s Principle which is also called the “Principle of Least Action”. The
Lagrangian formulation works with purely generalized coordinates, and generates
directly the equations of motion in the generalized coordinates. This save us the
trouble of identifying forces, which are vector quantities and transforming them
from rectangular Cartesian to any other coordinate system convenient for
studying the dynamics of the system.
In this Unit we first derive Lagrange’s equations of the second kind, using
D'Alembert’s Principle and the concept of generalized coordinates introduced
in Unit 2. We define the generalized velocity, generalized momentum and
generalized force for the dynamical system. We then introduce the Lagrangian
function and derive the Euler-Lagrange equations, which are equations of
motion derived from the Lagrangian function. You will see that for the vast
number of dynamical problems where the applied forces are conservative in
nature, the Lagrangian is merely the difference between the kinetic and
potential energies of the system and the Euler Lagrange (E-L) equations are
an elegant mechanism for generating easily a set of equations for all the
independent generalized coordinates Hence they are a complete solution to
the dynamical problem at hand. In Sec. 3.3 we extend the formalism for a
special class of generalized or velocity dependent potentials, for systems in
which the generalized force cannot be written as the gradient of a scalar
potential. In Sec. 3.4 you will study some applications of the Lagrangian
formulation. In particular, you will see how the E-L equations generate the
same equations of motion that one derives from Newton’s second law of
motion.
There are several examples, SAQs and TQs in this Unit and you must work
through these diligently to see for yourself the power of this formulation of
mechanics in solving dynamical problems. At the end of this Unit we will give
you an Appendix on the Calculus of Variations. You must study this appendix
before you start studying the next Unit, where we derive Lagrange’s equations
of motion from variational principles.
Expected Learning Outcomes
After studying this unit, you should be able to:
 derive expressions for the generalized velocity and generalized force in a
constrained system;
 derive Lagrange’s equation of motion of the second kind for an N-particle
system;
 obtain Lagrange’s equation of motion of the second kind for different
dynamical systems;
 derive the Euler-Lagrange(E-L) equation of motion for an N-particle
system; and
 obtain the Euler-Lagrange(E-L) equations of motion for different
70 dynamical systems.
Unit 3 Lagrange’s Equations and its Applications

3.2 LAGRANGE’S EQUATIONS FROM


D'ALEMBERT’S PRINCIPLE
3.2.1 Lagrange’s Equations of the Second Kind
Let us consider a N-particle dynamical system with holonomic constraints
expressed by equations of the form:
  
f (r1, r2 ,......, rN , t )  0,   1, 2, 3...., k (3.1)
Since there may be, in general an explicit time dependence, the constraint is
also rheonomic. If we have k (k < 3N) such equations of constraint then the
system has 3N  k degrees of freedom and can be described by the set of
3N  k independent generalized coordinates q1, q 2 ,.......q 3N  k . Let us say
  
that the the set of position vectors (r1, r2 ,......, rN ) is described by the set of
rectangular Cartesian coordinates  x1(t ), x 2 (t ), x 3 (t ),  x 4 (t ), x 5 (t ), x 6 (t ),...,

 x3N  2 (t ), x3N 1(t ), x3N (t ) where r1  x1(t ), x 2 (t ), x3 (t ) ,

r2   x 4 (t ), x 5 (t ), x 6 (t ) and so on. The following transformation equations
describe the relation between  x1(t ), x 2 (t ), x 3 (t ),  x 4 (t ), x 5 (t ), x 6 (t ),.....,
 x3N  2 (t ), x3N 1(t ), x3N (t ) and the set of independent generalized
coordinates q1, q 2 ,.......q 3N  k :
x1  x1q1, q2 ,.......q3N  k , t 
x 2  x 2 q1, q 2 ,.......q3N  k , t 

x3N  x3N q1, q2 ,.......q3N  k , t  (3.2)

Now, defining n  3N  k , we write the differential element dx i (i  1,2...,3N ) in


terms of dq i (i  1,2..., n ) as follows:

x i x x x
dx i  dq  i dq  ..... i dqn  i dt
q1 1 q 2 2 qn t
n
x i x
  q j
dq j   i dt
t
(3.3)
j 1

dx
The Cartesian velocities x i  i (i  1,2...,3N ) can then be written in terms of
dt
dq i
q i  (i  1,2..., n ) as:
dt
dx x dq1 x i dq 2 x dq n x i
x i  i  i   ..... i 
dt q1 dt q2 dt qn dt t
n
x i  x
  q j
qj  i
t
(3.4)
j 1
dq j
q j  is the total time derivative of the generalized coordinate q j and is
dt
called the generalized velocity.
Using the summation convention (sum over repeated indices) we can
dispense with the summation sign in Eq. (3.4) and write it simply as
x x
x i  i q j  i (3.5)
q j t 71
Block 1 The Lagrangian Formulation of Mechanics
where the index j is summed over from 1 to n. Because the Cartesian
coordinates xi are explicit functions only of the generalized coordinates qi
and time t and not of the generalized velocities q i , Eq. (3.5) also implies:
n
x i x i x
q l

 q j
 jl  i
q l
(3.6)
j 1

Eq. (3.6) is also called the dot cancellation equation. Next we evaluate :
d  x i 
  . Since x i are explicit functions only of the generalized coordinates
dt  q l 
x i
q i and time t, can be an explicit function of q j and time t as well, and so
q l
we can write:
 x    x i    x 
d  i    dq j    i dt (3.7)
 q l  q j  q l  t  q l 
and
d  x i    x i    x 
    q j   i  (3.8)
dt  q l  q j  q l  t  q l 

Where the summation convention, as usual implies a summation over the


repeated indices. We can also write Eq. (3.8) as:

d  x i    x i    x i 
  

dt  q l  q l  q q j  q  t 
 j l

  x i   x  x
  q j  i   i (3.9)
ql  q j  t  ql

In deriving Eq. (3.9) we have used the definition of x i from Eq. (3.5) and also
made use of the following relations:
  x i    x i    x i    x i 
    and     (3.10)
q j    q    q l  t 
q
 l q l  j  q l  t 

Note further that Eq. (3.9) is true, irrespective of whether the constraints in the
system are rheonomic or scleronomic, that is irrespective of whether x i
depends explicitly on t or not.
Now, since in a virtual displacement, the time is fixed (dt  0) , we can write
the virtual displament x i  dx i dt  0 . Likewise the virtual displacement in the
generalized coordinate is also q i  dq i dt  0 Substituting dt  0 in Eq. (3.3)
we get:
x
x i  i q j (3.11)
q j
Let us now go back to D Álembert’s Principle for an N-particle system, which is

 Fi A  p i . ri
N   
0 (3.12)
i 1
 
Fi A is the force on the ith particle and ri .

72
Unit 3 Lagrange’s Equations and its Applications
We can rewrite Eq.(3.12) in the 3N Cartersian coordinates for the N-particle
For example, m1 is
 d 2 xi
system { x i }, i  1,2....3N and using p i  mi  mi xi , where mi is the the mass of the
dt 2 particle associated
mass of the particle associated with the coordinate x i (see margin remark), with the
coordinates  x1, x 2 , x3 
we can write
, m2 is the mass of

 Fi 
3N
A the particle associated
 mi xi x i  0 (3.13) with the coordinates
i 1
 x4 , x5 , x6  and so on.
This is now a scalar equation, in which Fi A is the component of the force

along the coordinate x i ( whereas in Eq. 3.12, Fi A is the force on the ith
particle). Note that because the coordinates are not all independent due to the
constraints connecting them(as we have also explained in Unit 2), this cannot
be written as the set of 3N independent equations of the form:

F1A  mi x1   0 ; F2 A  mi x2  0 ;...... (3.14)


Writing x i in terms of qi we can write the D'Alembert’s principle in terms of
the virtual displacement of the generalized coordinates as

 Fi A  mi xi  q ij q j


 x 
0 (3.15)
i, j

The summation over j runs from 1 to n, where n  3N  k is the number of


generalized coordinates. Now let us rewrite the first term in Eq. (3.15) as:
 x  n
 Fi A  q i q   Q j q
 j j (3.16a)
i, j  j  j 1

Using the notation


 x 
Qj   Fi A  q ij  (3.16b)
i

where Q j can be looked upon as a generalized force along the jth


generalized coordinate. To the extent that the applied force is given by
some known force law, and we can write Fi A as a function of the Cartesian
coordinates and the velocities, Q j can also be written at the function of the
generalized coordinates and generalized velocities.
Notice that while x i , being the rectangular Cartesian coordinates, have the
dimensions of length, the generalized coordinates, q i , may not and therefore
Q j also, may not have the dimensions of force, and hence the term
generalized force. Remember however that Q j q j will always have the
dimensions of work.
Next we simplify the second term in the RHS of Eq. (3.15). Using Eq. (3.11)
we can write:
n  x 
mi xi x i   mi xi  q ij q j (3.17)
j 1
73
Block 1 The Lagrangian Formulation of Mechanics
Now, we can write:

d    x i   x i  d  x i 
mi x i     mi xi    mi x i  
dt   q   q  dt  q j 
 j   j  

 x   x 
 mi xi  i   mi x i  i
  q

 (3.18)
 q j   j 

d  x i  x i
Where we have used Eq. (3.9) to substitute  
dt  q j  q in Eq. (3.18). So
 j

 x  d   x i   x 
mi xi  i  
 dt mi x i

 q
  mi x i  i
  q

 (3.19)
 q j    j   j 
x i
Using the dot cancellation equation of Eq. (3.6) , we can substitute by
q j
x i
which gives us
q j
 x  d  x i   x 
mi xi  i     mi x i
 
  mi x i  i
  q

 (3.20)
 q j   dt  q j   j 
The total kinetic energy of the N-particle system, is:
1
T   2mi x i 2 (3.21)
i
where mi is the mass of the body with the coordinate x i . From Eq. (3.4) we
know that x i and hence T will be in general a function of generalized
coordinates, generalized velcities and time( q l , q l and t ). Therefore
T  T (q l , q l , t ) and we can write:
dT d  1  2  1   x i 

dq j dq j

 2 i i
mx 
 mi  2 x i 
q j 
 i  i 2 
x i
  mi x i
q j
(3.22a)
i
and
dT d  1  2  1   x i 

dq j dq j

 
 2 i i
mx 
  2
mi  2 x i
 q j



 i  i
x i x i
  mi x i
q j
  mi x i
q j
(3.22b)
i i
With this,
 x i  d  x i   x 
 mi xi 
 q
 j



  mi x i 
dt q j
  mi x i  i
  q


i i     j 
d  T  T
   (3.23)
dt  q j  q j
Substituting from Eqs. (3.16 and 3.23) into D'Alembert’s Principle (Eq. 3.13)
we get:

 Fi A  mi xi  q ij q j


 x 

74 i, j
Unit 3 Lagrange’s Equations and its Applications

 d  T  T 
  Q j  dt  q j   q j  q j  0 (3.24)
j
So far we have reformulated D'Alembert’s principle in terms of generalized
coordinates. In our derivation we have assumed the following:
i) The virtual work done by the forces of constraint is zero, and
ii) The constraints are holonomic, and qi are the set of generalized
(independent) coordinates that satisfy these constraints.
Since all the virtal displacements corresponding to the independent
generalized coordinates are independent, for Eq.(3.24) to be true, the
coefficient of each term should be zero and so:

d  T  T
Qj   
dt  q j  q  0 for all j (3.25)
 j

d  T  T
 
or
dt  q j  q  Q j (3.26)
 j

Notice that these are n  3N  k independent equations for the n  3N  k


generalized coordinates q1, q 2 ,.......q n . This is therefore a complete solution of
the dynamical problem and no extra conditions are required to determine the
solution. These equations are Lagrange’s equations of motion of the
second kind (1788).
Note that these equations contain no explicit dependence on the constraints
in the system. The constraints are implicitly taken care of in the definition of
the generalized coordinates and the transformation equations connecting the
generalized coordinates to the Cartesian coordinates.
Since each of these equations is a second order differential ,the general
solution of these equations contains 2n constants of integration. As always, if
the state of the system is known at any instant of time, i.e we know the values
of the generalized coordinates and generalized velocities at any instant of
time, we can uniquely determine these 2n constants of integration. And then
the state of the system is completely determined for all times.
Example 3.1
Derive the expression for the generalized force and Lagrange’s equation of
motion of the second kind for the systems:
a) A simple pendulum with a bob of mass m tied to a support by string of
constant length L0.
b) The Atwood’s machine consisting of of two blocks of wood of masses m1
and m2 , respectively, connected by an inextensible string of length l ,
assuming the radius of the pulley to be zero.
Solution : a) The Cartesian coordinates of the pendulum are ( x, y , z ) (where
we have used x1  x; x 2  y , x 3  z ) , the only applied force is in the direction
of x 2 , so in Eq. (3.15) F1A  0; F2A  mg; F3A  0 (Fig. 2.31)and there is just
one generalized coordinate is q1   . So:

x1  x  L0 sin  ; x2  y  L0 cos ; x3  z  0 75
Block 1 The Lagrangian Formulation of Mechanics
  x 
In Q j  
 A  x
Fi  i

 ,
The generalized force is Q j  Fi A  i
 q j
  (see margin remark) .


i
i  q j 
for each value of j ( j x x x x
runs from 1 to n , that is Q1  F1A 1  F2A 2  F3A 2  F2A 2
q1 q1 q1 q1
for each generalized
coordinate), this (L0 cos )
summation has to  mg  mgL0 sin  (i)

carried out over all i,
which is the number of To calculate the kinetic energy in terms of the generalized coordinate, we
Cartesian coordinates
write
in the dynamical
system, i.e. 3N. Here 1 1 1
T  mx12  mx 2 2  mx 3 2
the number of 2 2 2
Cartesian coordinates
1  2 1  2 1 2
is 3 and the number of  mx  my  mz
generalized coordinates 2 2 2
is 1. Since x  L0 cos   ; y  L0 sin   ; z  0

We get
1 1
T  mL0 2 (cos2   2  sin2   2 )  mL0 2  2 (ii)
2 2
d  T  T
For this system Eq. (3.26) is    Q1 .
dt  q1  q1

Substituting from Eqs. (i) and (ii) we get

d    mL 2  2     1 2 2
  0    mL0    mgL0 sin 
dt 
   2    2 
 
d
 (mL0 2  )  mgL0 sin   mL0 2   mgL0 sin 
dt
We get the following equation of motion

   g sin 
L0
b) Let us now label the coordinates of the masses by  x1, y 1, z1  and
 x 2 , y 2 , z2  . In the notation of Sec. 3.2.1 we have x1  x1; x 2  y 1; x 3  z1
and x 4  x 2 ; x 5  y 2 ; x 6  z2 . The applied forces are the gravitational
forces, which act only along the y-direction. There is just one generalized
coordinate which we choose as q1  y 1 and y 1  y 2  l (Fig. 2.15). The
generalized force is
6
 x  y y
Q1   F A q1i   m1g y11  m2 g y21
i
i 1

y1  l  y1
 m1g  m2g  m1g  m2g
y1 y1

The kinetic energy is


1 1 1
T  m1y12  m2 y 2 2  ( m1  m2 ) y12
2 2 2
76
Unit 3 Lagrange’s Equations and its Applications
Since y 2  l  y 1 we have y 2   y 1 . Lagrange’s equation of motion of the
second kind (Eq. 3.26) is:
d   1 2   1 2
  (m1  m2 ) y1    (m1  m2 ) y1   m1g  m2 g
dt  y 1  2
  y 1  2 
(m  m2 )g
which gives us y1  1
m1  m2

You may now like to work out an SAQ on this section.

SAQ 1
Derive the expression for the generalized force and Lagrange’s equation of
motion of the second kind for a block of mass m sliding down a frictionless
incline of mass M which in turn is sliding on a smooth horizontal plane.

3.2.2 Lagrangian
Let us now assume that all the applied forces in the system are conservative.

In that case, these forces are derivable from a scalar potential. If F is the
applied force on the kth particle in the system, then :
 
F   k V (3.27)

where V is a function of the coordinates only , so V  V { x i }  and  is the
gradient of V with respect to the position vector of the particle, which is, say

r  x iˆ  x ˆj  x kˆ :
k a b c
 V ˆ V ˆ V ˆ V V V
F  i  j k  Fa   ; Fb   ; Fc  
x a x b x c x a x b x c

V V
So Fa   is the force along the coordinate xa , Fb   is the force
x a x b
along the coordinate x b , and so on. Therefore in an N-particle system:
 V ˆ V ˆ V ˆ  V ˆ V ˆ V ˆ
F1   i  j k ; F2   i  j k ;... (3.28)
x1 x 2 x 3 x 4 x5 x 6
The generalized force (Eq. 3.16b) Q j is

 x   x   V  x 
Qj   Fi A  q ij     iV  q ij     x i  q ij 
i i i

V q i , t 
 (3.29)
q j

Note that while V is a function of x i  , but because our holonomic constraints


may be either rheonomic or scleronomic, x i  x i q i , t  (Eq. 3.2), hence
V  V x i   V q i , t  . What is implicit in Eq. (3.29) is the following:

When the forces in the system are a sum of conservative and constraint
forces(which do no virtual work) and the constraints in the system are
holonomic constraints, the generalized force can be derived from the
partial derivative of the potential energy written in terms of the
generalized coordinates. 77
Block 1 The Lagrangian Formulation of Mechanics
Substituting for Q j from Eq. (3.29) in Lagrangre’s equation of motion of the
second kind (Eq. 3.26), we get

d  T  T V
   (3.30)
 
dt  q j  q j
 q j

which is:

d  T  T  V 
  0 (3.31)
dt  q j 
 q j

Now, since V  V q i , t  is not a function of the generalized velocities q i  ,


we can write
V
0 (3.32)
q i

d  V 
Therefore we can add a term of the form   ( 0) to Eq. (3.31) to write
dt  q j 

d  T  V   T  V 
  0 (3.33)
dt  q j  q j
We now define a function of the generalized coordinates q i  , where the
generalized velocities q i  and time t, called the Lagrangian Lq i , q i , t  as
L  T V (3.34)
The Lagrangian is the difference between the kinetic and potential energies of
the dynamical system. With this Eq. (3.33) is written as:
d  L  L
   0 (3.35)
dt  q j  q j
There will be n such second order differential equations, one for each
generalized coordinate. This set of equations, which completely describe the
motion of the system are called the Euler – Lagrange Equations of
Motion(which is the name we will be using henceforth in this block). They are
also referred to as just Lagrange’s Equations of Motion.
It is important to remember that these equations are for a system with
holonomic constraints, where the applied forces are conservative and
derivable from a potential function of the kind V  V x i  .
At this stage for notational simplicity we drop the curly brackets in the symbol
for the Lagrangian to write Lq i , q i , t  as Lq i , q i , t  ,remembering, that in
general, the Lagrangian is a function of all generalized coordnates and
velocities. The Lagrangian for a mechanical system is not unique. In fact for
any Lagrangian Lqi , q i , t  for a system, we can define another Lagrangian
L q i , q i , t  with the help of an arbitrary function F q i , t  , where F q i , t  is
a differentiable function of the generalized coordinates, as follows:
dF
L q i , q i , t   Lq i , q i , t   (3.36)
dt
This new Lagrangian will give rise to the same set of equations of motion. Let
us prove this property in the following example. F q i , t  is also called the
78 guage function for the Lagrangian Lqi , q i , t  .
Unit 3 Lagrange’s Equations and its Applications

Example 3.2
Let Lq i , q i , t  be the Lagrangian of a system with n degrees of freedom,
where the generalized coordinates are q1, q 2 ,.......q n . If F q i , t  is any
arbitrary function of the generalized coordinates and time , which is
differentiable with respect to the generalized coordinates, show that a
Lagrangian L q i , q i , t  defined by
dF
L q i , q i , t   Lq i , q i , t  
dt
d  L   L 
Satisfies the set of equations:    0.
dt  q i  q i

Solution : We have
L   L  dF   L    dF 
     (i)
q i q i  dt  q i q i  dt 

and
L   dF  L   dF 
 L      (ii)
q i q i  dt  q i q i  dt 
F  F q i , t  we can write
dF F  F
dt
 
q i
qi 
t
i
which gives us
  F   F
  (iii)
q i  dt  q i
Now
d L  d  L  d    dF  
    
dt q i
 dt  q i  dt  q i  dt  
d  L  d  F 
     (iv)
dt  q i  dt  q i 
So we have
d L  L  d  L  d  F  L   dF 
         (v)
dt q i q i dt  q i  dt  q i  q i q i  dt 
 
Since the Lagrangian L satisfies the Euler-Lagrange equation, therefore
d  L  L
   0 . Further since F is a differentiable function we must have:
dt  q i  q i
d  F    dF 
   
dt  q i  q i  dt 
d L  L 
Eq. (v) then reduces to:   0.
dt q i q i
Hence L satisfies the Euler-Lagrange equations.

The set of generalized coordinates for the system is also not unique. The form
of the Euler-Lagrange equations remain the same for any set of
independent generalized coordinates. You can prove that for yourself in the
following SAQ. 79
Block 1 The Lagrangian Formulation of Mechanics

SAQ 2
A system with n degrees of freedom is described by the set of independent
generalized coordinates are q1, q 2 ,.......q n and obeys the Euler-Lagrange
equations (Eq. 3.35). Suppose that we now select another set of independent
coordinates u1,u 2 ,....... u n which are connected to the coordinates
q1, q 2 ,.......q n by the following set of transformation equations:

q1  q1u1, u2 ,.......un , t 

q2  q2 u1, u2 ,.......un , t 

 (3.37)
qn  q n u1, u2 ,.......un , t 

Rewriting the Lagrangian as a function of ui  , the generalized velocities ui 


and time t using the equations of transformation, show that:

d  L   L 
   0
dt  u j  u j
where L   Lq i u i , q i u i , t  which is the Lagrangian written as a function
of the new variables.

Such transformations, as described by the set of equations (Eq. 3.37) are


called point transformations. This result tells us that the form of the Euler-
Lagrange equations is invariant under a point transformation.
This tell us something important about the E-L equations. The form of the
Euler Lagrange equations of motion remain the same, no matter what be our
choice of generalized coordinates. Unlike this, the form of the equation of
motion derived from Newton’s laws is not the same in different coordinate
systems.
Example 3.3
Derive the Euler-Lagrange equations of motion for the following systems
a) A simple pendulum with a bob of mass m tied to a support by string of
constant length L0 .
b) The Atwood’s machine consisting of two blocks of wood of masses m1 and
m2 , respectively, connected by an inextensible string of length l, assuming
the radius of the pulley to be zero.
Solution : a) We have already written the kinetic energy for the simple
pendulum in terms of the generalized coordinate  in Example 3.1a

1
T  mL02  2
2
The potential energy (relative to the x-axis) is V  mgL0 cos  .

The Lagrangian is:


80
Unit 3 Lagrange’s Equations and its Applications
1
L()  m L0 2  2  mg L0 cos  (i)
2
The Euler-Lagrange (E-L) equation for the generalized coordinate , is

d  L  L
   0
dt    

From Eq. (i)


L L
 mL0 2  ;  mg L0 sin  (ii)

 

Substituting from Eq. (ii) in Eq. (i) we get


g
mL0 2   mgL0 sin   0     sin 
L0

b) From Example 3.1b, the kinetic energy of the system is


1
T  (m1  m2 )y 12
2
The potential energy is
V  m1gy 1  m 2 gy 2  m1gy 1  m 2 g (l  y 1 )  (m1  m2 )gy1  m2g l
So the Lagrangian is
1
L  T V  (m1  m2 )y12  (m1  m2 )gy1  m2g l
2
The E-L equation of motion is:
d  dL  L
   0  (m1  m2 )y1  (m1  m2 )g  0
dt  y1  y 1
(m1  m 2 )g
The equation of motion is y1  .
m1  m 2

You can now try to derive the equations of motion for some systems.

SAQ 3
Derive the Euler-Lagrange equations of motion for the following systems:
a) A bead of mass m sliding due to gravity on an elliptical wire.
b) A block of mass m sliding down a frictionless incline of mass M which in
turn is sliding on a smooth horizontal plane.

3.2.3 Kinetic Energy in Generalized Coordinates


Let us now write the kinetic energy of the system explicitly in terms of the
x i  x
generalized coordinates. Using the relation x i  q j  i , the kinetic
q j t
energy of the system is:

81
Block 1 The Lagrangian Formulation of Mechanics
2
1 2 1  x i  x 
T   2
mi x i  mi
2  j q j

qj  i 
t 

i i  

1  x 
2  x  x
  mi  i    mi   i q j  i
 q j  t
i 2  t  i  j 
1  x  x 
  m i   i q j   i q k  (3.38)
 q j  q k 
i 2  j  k 

 T0  T j q j  T jk q j q k
j j ,k
where
2
1 x
T0   2mi  ti  (3.39a)
i

 x  x i 
T j   mi  i   (3.39b)
 t  q j 
i 
1  x  x 
T jk   2mi  q ij  q ki  (3.39c)
i

Each of these functions can be derived from the transformation equations


connecting x i  and q i  . Notice that the first term in the expression for T, i.e.
T0 in Eq. (3.38) is independent of the generalized velocities, the second term
is linear in the generalized velocities and the third term is quadratic in the
generalized velocities. Since x i  are functions of q i  and t, T0 , T j and T jk
may be functions of the generalized coordinates and time.

However, if the holonomic constraint in the system is also scleronomic, the


transformation equations will not depend explicitly on time, hence:
x i
 0 for all i (3.40)
t

And we get: T  T jk q j q k (3.41)


j ,k

Which means that for a system with holonomic scleronomic constraints the
kinetic energy is a homogenous quadratic form in the generalized velocities.

We have derived the Euler-Lagrange equations of motion for a system in


which the applied force is derivable from a velocity independent potential. Let
us how the formalism holds when the potential is velocity dependent.

3.3 LAGRANGE’S EQUATIONS FOR A


VELOCITY-DEPENDENT POTENTIAL
Let us start with Eq. (3.26) that is the Lagrange’s equation of motion of the
d  T  T
second kind    Qj.
dt  q j  q j

82
Unit 3 Lagrange’s Equations and its Applications
Let us consider the case in which the generalized force cannot be written as
V q i , t 
Qj   . Instead let us assume that the generalized force is derived
q j
from a general potential function U  U q i ,q i , t  as follows:

U d  U 
Qj      (3.42)
q j dt  q j 

Substituting for Q j from Eq. (3.42) into Lagrange’s equation of motion of the
second kind, we get:

d  T  T U d  U 
   
dt  q j  q   q  dt  q  (3.43)
 j j  j 

d  T  U   T  U 
or   0 (3.44)
dt  q j  q j

Now defining the Lagrangian for the system as L  T  U we get back the
familiar form of the Euler-Lagrange equation, this time for a velocity
dependent potential U which statisfies Eq.(3.42).

The electromagnetic force field, as we know is derived from a velocity


dependent potential. In the following example we derive the equation of
motion for a charge moving in an region containing an electric and magnetic
field.
Example 3.4
Consider an electric chargeq of mass m moving in a  charge free region
containing an electric field E and a magnetic field B . Assume that the force
on the charge can be derived from a general potential function of the kind:
 
U  q  qA . v
 
where   ( x, y , z, t ) and A  A( x, y , z, t ) are the scalar and vector potentials
respectively, with

  A   
E     and B    A
t
Deduce the equations of motion for the charged particle.
Solution : The coordinates of the charged particle are (x,y,z ).The
Lagrangian for the system is
1  
L T U  mv 2  q  qA.v
2
1
 m( x 2  y 2  z 2 )  q  q( Ax x  Ay y  Az z ) (i)
2
From Eq. (i) we get
L d  L  dA
 mx  qAx     mx  q x ;
x
 dt  x 
 dt
83
Block 1 The Lagrangian Formulation of Mechanics
L   A Ay A 
 q  q  x x  y  z z 
x x  x y z 

The E–L equation for x is:


 A Ay A   d Ax 
mx  q  x x  y  z z   q    (ii)
 x x x   x dt 

The total time derivative of Ax is


dA x A x A A A A x  
  x x  y x  z x   v . A x (iii)
dt t x y z t

Using Eq. (iii) in Eq. (ii) we get:


 A Ay A 
mx  q  x x  y  z z 
 x x x 

  Ax  Ax  Ax  Ax 


 q  x y z 
 x t x y z 

 Ay  Ax  Az  Ax     Ax 


 q  y y z z   q 
 x y x z   x t 

 Ay Ax   A A    Ax 


 qy     qz  z  x   q    (iv)
 x y   x z   x t 

  A   
From E    and B    A we get:
t

  A   Ax  A
E x           qE x  q   x  (v)
 t  x x t  x t 

v  B x  v    Ax  y  Axy 


 A x A A
   z  z  x 
y   x z 
 

    Ay Ax    Az Ax


  
q v  B x  q  y     z  
 (vi)
  x y   x z 

Using Eqs. (v) and (vi) in Eq. (iv) we get the equation of motion in the x-
direction as
mx  qE x  (v  B ) x 

Let us now look at some applications of the Lagrangian formulation in simple


mechanical systems. Before that, let us emphasize the properties of the
Lagrangian.
1. The Lagrangian of the system L is a scalar function of the independent
generalized coordinates q i  , the generalized velocities q i  and time t .

2. The number of generalized coordinates is equlivalent to the number of


degrees of freedom of the system, in keeping with the existing constraints.
The set of variables qi  describes a point in an abstract n-dimensional
configuration space. The set q i  constitute the tangent direction of the
84 point. Together the set qi , q i describe the dynamical state of the system
Unit 3 Lagrange’s Equations and its Applications
at a time t . With the help of the Euler-Lagrange equations we determine
the evolution of the dynamical state.
3. The constraints on the system are holonomic. They can be however, either
scleronomous or rheonomous.
4. The Lagrangian is defined as L  T  V or L  T  U , depending on
whether the potential is velocity independent ( V  V q i , t  ) or velocity
dependent ( U  U q i ,q i , t  ) respectively. Whereas V  V q i , t  is the
potential energy of the system, U  U q i ,q i , t  cannot be described as
the potential energy in the conventional sense, it is a generalized potential
because it depends on more than the position of the particle (we cannot
derive it from the line integral of the generalized force as we can
V  V q i , t  ) . The velocity dependent potential U must satisfy Eq. (3.42)
for the form of the Euler-Lagrange equations to hold.

3.4 EULER-LAGRANGE EQUATIONS OF MOTION


FOR SIMPLE SYSTEMS
In this section we obtain the Euler- Lagrange equations of motion for some
simple systems.

3.4.1 Unconstrained Motion of a Single Particle


Since the motion is unconstrained, there are three degrees of freedom.
Therefore we need three independent generalized coordinates to describe the
motion. We work out this problem using both Cartesian coordinates and
spherical polar coordinates. Let us work first work out this problem using
Cartesian coordinates.

i) Rectangular Cartesian Coordinates

Let us consider the generalized coordinates to be the Cartesian


coordinates of the particle, i. e. q1  x; q 2  y ; q3  z . The generalized
velocities are q1  x; q 2  y ; q 3  z. The kinetic energy is:

T 
2

1  2  2 2
m x y z  (3.45)

Assuming that the potential energy is V  V x, y , z  , we get

1 2  2 2
L  T V  mx  y  z   V ( x, y , z ) (3.46)
2

We can write the following three E-L equations:

d  L  L d  L  L d L L
   0;    0 ;     0 (3.46)
dt  x  x dt 
 y y dt  z  z

where L  T  V . For the x coordinate we can write

d  (T  V )  (T  V )
  0 (3.47)
dt  x  x
85
Block 1 The Lagrangian Formulation of Mechanics
T V d (T  V )  d d
 mx;  0     (mx )  ( px ) (3.48)
x x dt  x  dt dt

(T )
where p x is the x-component of the momentum. Further  0 . So
x

dp x V dp x
 0  Fx (3.49a)
dt x dt

V
where Fx   is the x -component of the force. Similarly
x
dp y
 Fy (3.49b)
dt

dpz
and  Fz (3.49c)
dt

where p y is the y-component of the momentum, pz is the z-component


V V
of the momentum, Fy   and Fz   . So we get back the
y z
equations of motion as obtained from Newton’s second law.

ii) Free Particle in Spherical Polar Coordinates


Let us consider the generalized coordinates to be the spherical polar
coordinates of the particle, i. e. q1  r ; q 2   ; q3   . The generalized
velocities are q  r ; q   ; q   . To write down the kinetic energy for a
1 2 3
particle of mass m and coordinates (r , , ) , we first define (r , , ) in terms
of ( x, y , z ) :
x  r sin  cos  ; y  r sin  sin  ; z  r cos  (3.50)

x 2  r sin  cos   r  cos  cos   r  sin  sin 


2
(3.51a)

y 2  r sin  sin   r  cos  sin   r  sin  cos  


2
(3.51b)

z 2  r cos   r  sin 


2
(3.51c)
1 2
The Lagrangian is: L  m(r  r 2  2  r 2 sin 2  2 ) (3.52)
2
The equations of motion for this Lagrangian are the usual:
d L L

dt r r

 m r  m r  2  sin2   2  0  (3.53a)

d L L
  m r 2   2 m r r   m r 2 sin  cos   2  0 (3.53b)
dt  
d L L
  m r 2 sin 2    2m r sin 2  r   2 m r 2 sin  cos     0 (3.53c)
dt  

Example 3.5
Consider a simple pendulum of length L0 and mass m for which the point of
support has a vertical motion described by A sin t . Determine the Euler-
86 Lagrange equation of motion for the pendulum.
Unit 3 Lagrange’s Equations and its Applications
Solution : The generalized coordinate for the simple pendulum is 
(Example.3.3a). We have earlier evaluated for the x coordinate:
x  L0 sin   x  L0 cos  

Since the point of support is oscillating, the y coordinate is now:


y  A sin t  L0 cos   y  A cos t  L0 sin  

The kinetic energy is:


1 1
T  mL0 2(cos 2   sin 2 )  2  mA 2  2 cos 2 t  m AL0 cos t sin  
2 2
1 1
 mL0 2  2  m a 2  2 cos 2 t  AmL0 cos t sin  
2 2
The potential energy is: V  mg ( A sin t  L0 cos )
1 1
L mL2  2  m A 2 2 cos 2 t  AmL0 cos t sin    mg ( A sin t  L0 cos )
2 2
dL L
 mL0 2   Am L0 cos t sin  ;  A m L0 cos t cos    mgL0 sin 

d 
d  L  2 2
    mL0   A mL0 sin t sin   A m L0 cos t cos  
dt   

The E-L equation of motion for  is


d  L  L
    0  mL0 2  A2 mL0 sin t sin   mgL0 sin   0
dt    

It is important to remember that for a system with holonomic constraints, the


number of generalized independent coordinates is equal to the number of
degrees of freedom of the system. Therefore, if you have an N-particle system
and k holonomic constraints, the system has n  3N  k degrees of freedom
and that many independent generalized coordinates. However if there are,
say, in addition to these k holonomic constraints,k1nonholonomic constraints
as well, the number of degrees of freedom of the system is n1  3N  k  k1 ,
but the number of generalized coordinates required is still n  3N  k . So in
general, in the presence of nonholonomic constraints the number of
generalized coordinates is more than the number of degrees of freedom of the
system.
Let us now summarize what you have studied in this Unit.

3.5 SUMMARY
 Generalized Velocity and Generalized Force
An N-particle dynamical system with k ( k  3N ) holonomic constraints
has n  3N  k degrees of freedom and can be described by the set of
n independent generalized coordinates q1, q 2 ,.......q n .

dq
The generalized velocity is q i  i (i  1,2..., n )
dt

87
Block 1 The Lagrangian Formulation of Mechanics
And the generalized force along the generalized coordinate q j is
3N  x 
Qj   Fi A  q ij  ( i  1,2...,3N and j  1,2..., n ) .
i 1

 Lagrange’s Equation of Motion of the Second Kind


For a dynamical system in which
i) The virtual work done by the forces of constraint is zero, and
ii) The constraints are holonomic, and q i (i  1,2...n ) are the set of
independent generalized coordinates that satisfy these constraints.
Starting from D'Alembert’s principle, we can derive n equations of
motion for the n independent generalized coordinates, which are
called the Lagrange’s equations of motion of the second kind:

d  T  T
   Qi ( i  1,2..., n )
dt  q i  q i

 Lagrangian of a Dynamical System

For a dynamical system with holonomic constraints, where the applied


forces are conservative and derivable from a potential function of the
kind V  V x i  , we define a function of the generalized coordinates,
generalized velocities and time, called the Lagrangian L(q i ,q i , t  of
the dynamical system as the difference between its kinetic and
potential energies:

L  T V

 Euler-Lagrange Equations of Motion


For a dynamical system with the Lagrangian defined as L(q i ,q i , t  ,
the equation of motion for each generalized coordinate is given by the
Euler-Lagrange (E-L) equation of motion:

d  L  L
   0 ( i  1,2..., n )
dt  q i  q i

 Kinetic Energy in Generalized Coordinates

The kinetic energy of a dynamical system with n degrees of freedom


can be written as:
1
T  2mi x i 2  T0  T j q j  T jk q j q k
i j j ,k

where
2
1  x i  1  x i  x i 
T0   2
mi 
 t
 ;T j 


2
mi 
 t

 q j
;


i i

1  x  x 
T jk   2mi  q ij  qki 
i

88
Unit 3 Lagrange’s Equations and its Applications
T0 is independent of the generalized velocities, the second term is
linear in the generalized velocities and the third term is quadratic in the
generalized velocities.

 Euler-Lagrange Equations for a Velocity Dependent Potential

If the generalized force in the dynamical system is derived from a


general potential function U  U q i ,q i , t  which satisfies the
following relation for the generalized force:

U d  U 
Qj     

q j dt  q j 

Then, defining the Lagrangian for the system as L  T  U , we can get


back the familiar form of the Euler-Lagrange equation, this time for a
velocity dependent potential U.

3.6 TERMINAL QUESTIONS


1. Obtain Lagrange’s equation of motion of the second kind for a bead of
mass m moving on a fixed elliptical wire.

2. A double pendulum is made up of two simple pendulums in which a lower


pendulum of mass mB and length LB is suspended from a upper
pendulum of mass mA and length LA . Obtain the Euler Lagrange
equations of motion for the system in which m A  mB  m (say).

3. Obtain the Euler-Lagrange equations of motion for a spherical pendulum


with a bob of mass m and length L0 .

4. A bead of mass m moves without friction along a fixed wire bent in the
shape of a parabola in the xy-plane with y  x 2 . Obtain the Euler-
Lagrange equation of motion for the bead.

5. A bead of mass m moves without friction along a straight wire. The wire is
is rotating about the y- axis with an angular speed  , at a fixed angle of
0 with the y-axis. Obtain the Euler- Lagrange equation of motion for the
bead.

6. A mechanical system with two degrees of freedom is described by the two


generalized coordinates u1 and u 2 . If the kinetic energy of the system is
T  u 2 2u12  2u 2 2 and potential energy is V  u12  u 2 2 , write down the
Lagrangian for the system. Derive the equations of motion.

7. Derive the Euler-Lagrange equations of motion for the Lagrangian

1 2 1
L mx  k ( x  a )2
2 2

8. Given the following Lagrangian for a particle of mass M and generalized


coordinate x:
89
Block 1 The Lagrangian Formulation of Mechanics

L  x 2 sin 2 t  xx sin 2t  x 2  2 


M
2

determine the E-L equations of motion.

3.7 SOLUTIONS AND ANSWERS


Self-Assessment Questions
1. The Cartesian coordinates are ( x1, y 1, z1 ) and ( x 2 , y 2 , z2 ) for the block and
the centroid of the moving plane, respectively. The two generalized
coordinates are q1  s and q 2  X (SAQ 5b and Example 2.6b, Unit 2).
We have for the block:
y

s

a

X

x
b
z

Fig. 3.1: Block sliding down a moving incline.

x1  X  s cos   x1  X  s cos  ;


y1  a  s sin   y1  s sin  z1  0

And assuming that the centroid of the inclined plane lies in the xy-plane,
we can write
x2  X  x0 ; y 2  y 0 ; z2  00  x 2  X ; y 2  z2  0 .

The kinetic energy of the system is


1 2 1 2 1 2
T  m x1  my 1  M x 2
2 2 2
1 2 1 1
 MX  m( X 2  s 2 cos 2   2 X s cos  )  ms 2 sin 2 
2 2 2
1
 M  m X 2  1 ms 2  mX s cos  (ii)
2 2
The only forces are the gravitational forces along the y-axis, so the
components of the force along the x and z coordinates of the block and the
plane are all zero. There is a force Fblock  mg along the y-coordinate of
the block that is y1 and a force Fincline  Mg along the y-coordinate of
the incline that is y 2 .
 The generalized force along s is
y 1 
Q1  Fblock  mg (a  s sin  )   mg sin 
q1 s

90
The generalized force along X is
Unit 3 Lagrange’s Equations and its Applications
y 2 y
Q2  Fincline  Fincline 2  0
q2 X

The equations of motion are


d  T  T d T   T
   Q1     Q1
dt  q1  q1 dt  s  s
 m scos 2   m X  cos   m s sin 2   mg sin 
 cos   mg sin   s  g sin   X
 m s  mX  cos 

And
d  T  T d  T  T
   Q2    0
dt  q 2  q 2 dt  X  X

mscos 
 M  m  X  mscos   0  X  
mM

L L qi L q i  u j u j 


2.   
 t  0  t 

u j i qi u j qi u j  
q i
Given that q i  q i (u1...u n , t ) ,  0 and
u j
q q q q
qi   i u j  i  i  i
j u j t u j uj

L  L q i L q i  L q i L q i
Further
u j
   qi u

 j q i u j    q i u j



q i u j
i i i

d  L   d  L  qi L  d  qi 


So        
dt  u j  i  dt  qi  u j  
i qI  dt  u j 


d  L  L d  L  qi L d  qi  L qi L q i 

dt  u j

 u        
 q i u j q i u j 
 j i  dt  qi  u j q i dt  u j  
 d  L  L  qi L  q i  L d  qi 
           
  u  u j  q i dt  u j 
i  dt  qi  qi 
 qi
 j    

The first term in the bracket is zero (E-L equation), so we get

d  L  L  d  q i  q i  L  q i q i 



dt  u j

 u 

  
dt u j

 u  q    
u j u j 
0
j i    j  i i 

3. a) The kinetic energy for the bead on the elliptical wire in terms of the
generalized coordinate  (as also calculated in TQ 1) is
1
T  m(a 2 sin 2   b 2 cos 2 ) 2
2
and the potential energy of the bead is V  mg b sin  .

1
The Lagrangian is: L  T  V  m(a 2 sin 2  b 2 cos2 ) 2  mgb sin 
2
L
 m(a 2 sin 2   b 2 cos 2 )
 91
Block 1 The Lagrangian Formulation of Mechanics
d  L 
    m(2a 2 sin  cos   2b 2 cos  sin ) 2
dt   
 m(a 2 sin 2  b 2 cos2 )

L 1
 m(2a 2 sin  cos   2b 2 cos  sin )  2  mgb cos 
 2
The E-L equation of motion for  is:
m(a 2 sin 2   b 2 cos 2 )  m(a 2  b 2 ) sin  cos   2
 mg b cos 

b) Using the results from SAQ 1 we can write the kinetic energy as
1 1
T  M  m  X 2   ms 2  mX s cos 
2 2
The potential energy is V  mg (a  s sin  )

1
 L  T V  M  m  X 2  1 ms 2
2 2

 mX s cos   mg s sin   mga.

The equations of motion are


d  L  L d  L  L
    0 and   0
dt  X  X dt  s  s
L L
 M  m  X  ms cos  ; 0

X X
L L
 ms  mX cos  ;  mg sin 
s
 s
The E-L equations are:
  mscos   0
(M  m ) X
 cos   mg sin   0  s  g sin   X
And ms  mX  cos .

Terminal Questions
y 1. The wire is fixed in the xy-plane and the Cartesian coordinates of the ball
are  x, y , z  as shown in Fig. 3.2. The constraints on the motion of the
( x, y , z )
bead are
b 
x x2 y 2
a   1; z  0
a2 b2
)
z There is only one generalized coordinates, say q1   . So
Fig. 3.2: TQ 1.
x x  a cos   x  a sin  ; y  b sin   y  b cos   ; z  0

The only force is the gravitational force and it only has a component
Fy  mg along the y-coordinate of the bead, so the generalized force is:

y (b sin )
Q1  Fy  mg  mgb cos  (i)
 

92 The kinetic energy is


Unit 3 Lagrange’s Equations and its Applications

T 
2

1 2  2 2 1

m x  y  z  m(a 2 sin 2   b 2 cos 2 ) 2
2
(ii)

T
So  m(a 2 sin   b 2 cos 2 )  (iii)


And
d
dt
 T  d
  
   dt

(ma 2 sin2   mb 2 cos 2 ) 
 m(a 2 sin2   b 2 cos 2 )  m (2a 2sin  cos   2b 2 cos  sin ) 2

 m(a 2 sin 2   b 2 cos 2 )   m (a 2  b 2 ) sin 2  2 (iv)


T 1 1 O
 m(2a 2 sin  cos   2b 2 cos  sin )  2  m (a 2  b 2 ) sin 2  2 (v) A
x
 2 2
LA
yA
Substituting from Eqs. (i), (iii), (iv) and (v) into Lagrange’s equation of motion
yB
of the second kind (Eq. 3.26 with q1  ) we get:
xA
A
m (a 2 sin2   b 2 cos 2 )  m(a 2  b 2 ) sin  cos   2  mgb cos  B LB
B
Or (a 2 sin 2   b 2 cos 2 )  (a 2  b 2 ) sin  cos   2  gb cos 
xB
2. The generalized coordinates are  A and B (Fig. 3.3, also Example 2.4).

For the first bob: y

x A  L A sin  A , y A  L A cos  A ; z A  0 Fig. 3.3: Double


Pendulum.
and x A  LA  A cos  A , y A  LA  A sin  A ; z A  0

The kinetic energy of the first bob is : TA 


2

1 2 2 1

m x A  y A  mL2A  2A
2

For the second bob:


x B  LA sin  A  LB sin  B , y B  L A cos  A  LB cos B ; zB  0

And
x B  LA  A cos  A  LB  B cos B ; y B  LA  A sin  A  LB  B sin B ; zB  0

The kinetic energy of the lower bob is

TB 
1
2
  1

m x B2  y B2  m L2A  2A  2LA LB  A  B cos(  B   A )  L2B  B
2
2

The total potential energy is
V  mgL A cos  A  mg (LA cos  A  LB cos B )

The Lagrangian is : L  TA  TB  V


1
2

m L2A  2A  L2A  2A  2LA LB  A  B cos( B   A )  L2B  B
2 
 mgL A cos  A  mg (LA cos  A  LB cos B )
1

 m 2L2A  2A  2LA LB  A  B cos(  B   A )  L2B  B
2
2 
 2mgLA cos  A  mgLB cos B (i) 93
Block 1 The Lagrangian Formulation of Mechanics
The E-L equations of motion are:
d  L  L d  L  L
   0;   0
dt   A   A dt   B   B
From Eq. (i) we get:
L
 2mL2A  A  mL A LB  B cos( B   A )
 A

d  L 
   2mL2A  A  mL A LB B cos(  B   A )
dt   A 
 mLALB  B  B   A  sin( B   A ) (ii)
L
 mL A LB  A  B sin(  B   A )  2mgL A sin  A (iii)
 A
L
 mL2B  B  mLA LB  A cos( B   A )
 B
d  L 
   mL2B B  mLA LB  A cos(  B   A )
dt   B 
 mLALB  A  B   A  sin( B   A ) (iv)
L
 mLA LB  A  B sin( B   A )  mgLB sin  B (v)
 B
We get the following E-L equations for  A and B :
2
2mL2A A  mL A LB B cos(  B   A )  mLA LB  B sin( B   A )  2mgL A sin  A

And
mL2B B  mLALB  A cos( B   A )  mLALB  A 2 sin( B   A )  mgLB sin B

3. The generalized coordinates for the spherical pendulum are  and  .


Using Eqs. (3.50 to 3.51c) for spherical polar coordinates with
1
r  L0  r  0 we get for the kinetic energy: T  mL0 2 ( 2   2 sin 2 )
2
The potential energy is : V  mg L0 cos  (assuming that the pendulum is
in the lower half).
1
The Lagrangian is: L  T  V  m L0 2 ( 2   2 sin2 )  mg L0 cos  (i)
2
From Eq. (i) we get:
L d  L   L  2
 mL0 2  ; 2 2
    mL0  ;    mL0 sin  cos    mg L0 sin 

 dt      
(ii)
And
L d  L  L
 mL0 2 sin2  ;    mL0 2 sin2   mL0 2 sin 2  ; 0 (iii)

 
dt    

The E-L equations are

94
Unit 3 Lagrange’s Equations and its Applications
d  L  L
    0  mL0 2  mL0 2 sin  cos   2  mgL0 sin   0
dt    

Or:   sin  cos   2  g sin 


L0

d  L  L
and     0   sin2    sin 2  
dt  
 

and from Eq. (iii) mL0 2 sin2   constant .

4. The coordinates of the bead are ( x, y , z ) as shown in Fig. 3.4.The


constraints on the motion of the bead are y  x 2 ; z  0 , since the wave is
fixed in the xy-plane. Bead has just one degree of freedom and therefore
just one generalized coordinate q1, which we can choose as the x-
coordinate of the bead . The kinetic energy of the bead is:
1
T  m( x 2  y 2 )
2 y

Since y 
d 2
dt
  1
x  2 xx we get : T  mx 2 (1  4 x 2 )
2
( x, y , z )
The potential energy (relative to the x-axis) is: V  mgy  mgx 2

1
The Lagrangian is: L  T  V  m(1  4 x 2 )x 2  mg x 2 (i)
2 x

d  dL  L
The E-L equation of motion is:   0 (ii) z
dt  x  x
Fig. 3.4: TQ 4.
From Eq. (i) we get:
L d  L 
 mx  4mx 2 x ;    m(1  4 x 2 )x  8mxx 2 (iii)
x
 dt  x 
z
L
 4mxx 2  2mgx (iv) 
x
Substituting from Eqs. (iii) and (iv) into Eq. (ii) we get the equation of ( x, y , z )
motion: (1  4 x 2 )x  4 xx 2  2gx 0

5. The constraints on the motion of the bead are (Fig. 3.5): O


t y
x2  y 2 y
 tan 0 ;  tan( t )
z x x
Fig. 3.5: TQ 5.
So the bead has just one degree of freedom which we choose to be we
distance r of the bead from the end the wire, O. The coordinates of the
bead are
x  r sin 0 cos t ; y  r sin 0 sin t ; z  r cos 0

The kinetic energy of bead is

T 
1
2
 1 1
m x 2  y 2  z 2  m r2  m2r 2 sin2 0
2 2 95
Block 1 The Lagrangian Formulation of Mechanics
Which is the sum of the kinetic energy due to its motion of the bead along
the wire and that due to the rotation of wire about the y-axis.
The potential energy is: V  mgr cos 0

1 2
The Lagrangian is L  m(r   2 r 2 sin 2 0 )  mgr cos 0 (i)
2
d  L  L
The Euler-Lagrange equation is   0 (ii)
dt  r  r
From Eq. (i) :
L d L L
 mr;    mr ;  m2r sin2 0  mg cos 0 (iii)
r dt  r  r
Substituting from Eq. (iii) into Eq. (i) we get the equation of motion for r as

mr  m2r sin2 0  mg cos 0 or r   2 r sin 2 0  g cos 0

6. The Lagrangian is L  u 22 u12  2u 22  u12  u 22 (i)

L L
So  2u1u 22 ;  4 u 2
u1
 u 2
 L  d L 
  2u1u22  4 u2 u1u2 ; 
d
   4 u2
dt  u1  dt  u 2 
L L
 2u1 ;  2u 2u12  2u 2
u1 u 2

The E-L equations of motion are:


d  L  L
    0  2u1 u 22  4 u 2 u1u 2  2u1  0
dt  u1  u1

d  L  L
And     0  4 u2  2u2u12  2u2  0
dt u
 2 u 2

L d L L
7. We have:  mx     mx ;  k ( x  a )
x dt  x  x
The equation of motion is mx  k ( x  a )  0

8. L 
M 2
2
x sin2 t  xx sin 2t  x 2  2 
L Mx
 M x sin2 t  sin 2t
x
 2
d  L  2 Mx
   M x sin t  2M x sin t cos t  sin 2t  Mx cos 2t
dt  x  2
L M 
 x sin 2t  Mx  2
x 2

The equation of motion is: sin 2 t x   sin 2 t x   x cos 2 t  x 2  0

96
Unit 3 Lagrange’s Equations and its Applications

APPENDIX 3A CALCULUS OF VARIATIONS

In 1744, Leonard Euler in his work The method of finding plane curves that
show some property of maximum and minimum introduced a general
mathematical method for the investigation of variational problems. In the
process he also formulated his version of the principle of least action, which is
the variational formulation of mechanics. This principle is significant not only
for analytical mechanics but also for modern theoretical physics including
general relativity and quantum mechanics.

Here we state the solution for the following purely mathematical problem
considered by Euler:
dy
Consider a known function F  y , , x  of three variables, the
 dx 
independent variable x, the dependent variable y (x ) and its first
dy
derivative with respect to the independent variable. The problem is to
dx
determine the curve y  y (x ) which will make the definite integral I over
x 0  y  x1 :

x1
 dy 
I[y ]   F  y, dx , x  dx (A3.1)
x0

an extremum.

Since the value of I [y ] would depend on the choice of the curve y (x ) , I [y ] is


also a functional of y. For the integral to be an extremum, there must be a
unique curve y (x ) for which I [y ] is either a maximum or a minimum. The
problem is analogous to the familiar statement in calculus, that when the
tangent to a curve is zero at a point, the point is either a maximum or a
minimum.
The solution to the stated problem exists only if the value of y (x ) is fixed at the
endpoints x 0 and x1 . There could be an infinite number of curves connecting
the two fixed end points, the problem then is to find the unique curve
y (x ) connecting the end points for which the integral I [y ] is a maximum or a
minimum.
Euler showed that the unique function for which the integral I [y ] is an
extremum is the solution of the following differential equation for y (x ) :

 
d  F  F
  0 (A3.2)
dx   dy   y
 
  dx  

This differential equation involves partial derivatives of F with respect to both


dy
y and . This equation is also referred to as the Euler Lagrange equation.
dx
97
Block 1 The Lagrangian Formulation of Mechanics
Let us now work out a simple problem to understand this equation.
Example
Determine the curve in the xy-plane that minimizes the length between two
points (0,0) and (a,b).
Solution : If ds is an infinitesimal distance along the tangent to the curve
between the two points in the xy-plane, then we can write:

ds 2  dx 2  dy 2  ds  dx 2  dy 2
Choosing x to be the independent variable, we can write y (a )  b . The
distance between the two points is
b 2
dy 
s[ y ]   1    dx
 dx 
0

The function y (x ) that gives the minimum distance between the points is the
extremum of this integral.

dy 2
Using F  1  in Eq. (A3.2) with
dx
dy
F F dx
 0, 
y dy dy 2
 1
dx dx
we get
dy
 0  y ( x )  cx  d
dx
The desired curve is therefore a straight line. The constants c and d can be
determined using the condition on the end points y (0)  0 and y (a )  b.

Since the choice of end points can be arbitrary, what is proved is that the
extremum distance between any two points is a straight line. In this case, the
extremum must be a minimum.

98
Unit 4 Hamilton’s Principle

UNIT 4
HAMILTON’S
PRINCIPLE
Structure

4.1 Introduction 4.3 Conservation Theorems


Expected Learning Outcomes Cyclic Coordinates and
4.2 Hamilton’s Principle and the Conservation of Generalized
Euler-Lagrange Equations Momentum
Configuration Space Conservation of the Energy Function
Hamilton’s Principle Integrals of Motion
Derivation of the Euler-Lagrange 4.4 Symmetry: Homogeneity and
Equations of Motion from Isotropy
Hamilton’s Principle 4.5 Hamilton’s Principle for
Generalized Momentum Nonholonomic Systems
Energy Function 4.6 Summary
4.7 Terminal Questions
4.8 Solutions and Answers
4.1 INTRODUCTION
In Unit 3 you have seen how using the Euler-Lagrange equations of motion
you can arrive at a complete solution of a dynamical problem by determining
its configuration at all times. The Euler-Lagrange equations, like the equations
of motion derived from Newton’s second law, are also second order differential
equations. The formalism holds for the vast class of dynamical systems that
are commonly studied in nature, which are defined by purely holonomic
constraints. It holds for normal potentials that depend only on the coordinates
of the system (V), as well as for velocity dependent potentials (U). So, the
Lagrangian is the scalar function L  T  V or L  T  U , as the case may
be, and the only condition is that T ,V / U must be first written down in the
inertial frame in which Newton’s second law is valid. Although a detailed
discussion on this is beyond the scope of the syllabus, the Lagrangian
formulation can also be extended to include certain dissipative forces
described by the Rayleigh dissipation function and for nonholomic constraints.
The Lagrangian is also used to define the action, another fundamental
variable that you will study in this Unit along with the “Principle of Least
Action”.
We have already derived the Euler-Lagrange Equations starting from the
D’Alembert’s Principle. However D′Alembert’s Principle, in its very statement
 
assumes Newton’s second law ( Fi A  p i ) with the important caveat that the
rate of change of momentum is determined only by non-constraint forces. 99
Block 1 The Lagrangian Formulation of Mechanics
Further the equations are derived from, what can be called a “differential”
A differential
principle. It is also possible to formulate dynamical problems in a whole new
principle, like the
D'Alembert’s Principle, way, independent of Newton’s law, using the “Variational Principle” which is
looks at the an “integral” principle (see margin remark) based on the calculus of
instantaneous state of variations, using Eq. (3A.2) of Appendix 3A.
motion of the system
In Sec. 4.2, we first introduce the idea of an n-dimensional configuration space
and considers small
virtual displacements for a dynamical system with n degrees of freedom, in which each point is
about that state. An labelled by n generalized coordinates. The evolution of the system with time is
integral principle represented by a curve in configuration space. We then introduce the action of
considers the entire a dynamical system and state the Hamilton’s principle, or the Principle of
motion of the particle Least Action, as it is often called. Using Euler’s equation we derive the Euler
between two instants of
Lagrange equations of motion for the system from the Hamilton’s Principle.
time say t1 and t 2 and
We also introduce the energy function and the notion of generalized momenta.
considers small virtual
In Section 4.3 we discuss the conservation theorems which follow from the E-L
variations of the motion
from the actual motion. equations of motion. In Sec. 4.4 we discuss the relation between the
conservations and the symmetries from which they follow. In Sec. 4.5 we
discuss Hamilton’s principle for nonholonomic systems.
Expected Learning Outcomes
After studying this unit, you should be able to:
 determine the configuration space for a dynamical system;
 obtain the generalized momenta and energy function;
 state Hamilton’s Principle and derive the Euler-Lagrange equations of
motion from Hamilton’s Principle; and
 state Hamilton’s Principle for a system with nonholonomic constraints and
derive the Euler-Lagrange equations of motion for it.

4.2 HAMILTON’S PRINCIPLE AND THE EULER-


LAGRANGE EQUATIONS
When I was in high school, my physics teacher called me down one
day after class and said, “You look bored, I want to tell you something
interesting”. Then he told me something I have always found
fascinating. Every time the subject comes up I work on it.
Richard Feynman on the “Principle of Least Action”
4.2.1 Configuration Space
Let us briefly recapitulate what the E-L equations of motion can give us for a
dynamical system (within the limitations under which the equations have been
derived). We are all familiar with describing motion in the 3-dimensional
Euclidean space, the so called “real” space. We imagine particles tracing
actual paths in this space where the position of each particle is described by a
position vector. In Unit 3, we moved away from this description of a system in
terms of its position coordinates to a set of generalized coordinates, which, in
general, cannot be split into sets of three coordinates, each set describing the
position of each particle. There is off course a prescription (the transformation
equations) that enable you to go back to the description of motion in real
space. However we now describe what is called a configuration space for the
100 system.
Unit 4 Hamilton’s Principle
Consider an N-particle dynamical system with n degrees of freedom which are
described by n independent generalized coordinates. The configuration of the
system at any instant of time is given uniquely by the set of values of these n
generalized coordinates at that instant of time. These n generalized
coordinates thus define an n-dimensional configuration space in which each
point is labelled by the set of n generalized coordinates, q1, q 2 ,.......q n . Each
coordinate has a physical meaning and the range of that coordinate is
determined by its physical meaning. For example if qi is an angle, it could vary
between, say 0 to 180, etc. At any instant of time, each point of the
configuration space is a possible configuration of the dynamical system. As
the system evolves with time it traces out a path in configuration space. An
infinite number of curves, each signifying a possible path of the system passes
through each point in configuration space, the actual path would be decided
by the initial conditions.

Let us say that we specify the configuration of the system at an instant of time,
 
say t  t 0 , i.e at t  t 0 the system is at the point P0 q10 , q 2 0 ,.......q n 0 in the
configuration space C. If, in addition, we also specify the values of the
 
generalized velocities q i at this instant, let us say that q10 , q 2 0 ,.......q n 0 are
 
the generalized velocities at P0 q10 , q 2 0 ,.......q n 0 at t  t 0 , then using the E-
L equations we can determine the configuration of the system for all later
times. The set of points in configuration space corresponding to the
dynamical evolution of the system with time, determined using the E-L
equations starting from an initial configuration and an initial set of generalzed
velocities, define a unique trajectory for the system in configuration space.

Alternatively, we could choose an initial configuration given by the point


 
P1 q11, q 21,.......q n1 at a time t  t1 and a final configuration which is point
P2 q12 , q 2 2 ,.......q n 2  at a time t  t 2 , then also the E-L equations will help us
(in general) find the unique trajectory for the system corresponding to its
dynamical evolution from P1 to P2 .

In both these approaches, the generalized coordinates and the generalized


velocities are determined at each instant of time. They are merely different
approaches in determining the integration constants in the solutions of the E-L
equations, which are second order differential equations.

So, solving the E-L equations is all about determining the unique trajectory for
the time evolution of the system in configuration space using the given initial
conditions. Let us now see how the second of these conditions coupled with
an important variational principle, can also determine the unique trajectory of
the system.

SAQ 1
Define the configuaration space for the following systems:
i) A spherical pendulum(TQ 3, Unit 3)
ii) An Atwood’s Machine(Example 3.3b, Unit 3)
101
Block 1 The Lagrangian Formulation of Mechanics
4.2.2 Hamilton’s Principle
Let us consider any two points in configuration space P1 q11, q21,.......qn1 at a  
time t  t1 and a final configuration 
P2 q12 , q2 2 ,.......qn 2 
at a time t  t 2 . We
can imagine that our system starts from from the point P1 at a time t  t1 and
reaches the point P2 at a time t  t 2 , following a path, say C between P1 and
P2 . The system would be represented by a particular point on the curve C at
a specific time, so at each point on the path, the generalized coordinates qi
would have definite values which depend on the time t, so we can say that the
generalized coordinates are functions of time, qi  qi (t ) , such that
corresponding to the end points of the path we have:
q i (t1 )  q i 1 and q i (t 2 )  q i 2 .
dq (t )
Similarly the generalized velocities defined by q i  i are also uniquely
dt
determined at each point on the path, and therefore, the Lagrangian Lqi , q i , t 
also has a definite numerical value at each point of the path C. Clearly the
value would depend on the path chosen between the points P1 and P2 . Thus
we can say that the state of the mechanical system along the path C is
characterized by the Lagrangian function Lqi , q i , t  , which is a definite
function of generalized cordinates, generalized velocities and time.
Then we can define a functional (which is a function of functions) called the
action S( sometimes referred to as the action integral) , for any physical
system along any path traversed by the physical system, between two points
in configuration space, as the line integral of the Lagrangian function along the
path as:
t2
S   Lqi , q i , t dt (4.1)
t1

S is a functional of the functions qi and Lq i , q i , t  , because it depends on the


functional forms of these functions expressed in terms of t. Clearly there are
an infinite number of possible paths between these two points and the value
of the action integral would be different for each of these paths.
Let us assume that the function Lq i , q i , t  is known for the physical system,
though the function qi (t ) is not known. Clearly stated, the problem is the
following, given a functional S of a known function of coordinates, velocities
and time, L, is it possible to determine q i (t ) ? It is, using the principles of
calculus of variations. (Recollect the statement of the Euler equation problem)
We can insert a trial function qi (t ) into Eq. (4.1) and determine the value of
the action integral for each such trial function. Each trial function qi (t ) defines
a path between the points between P1 and P2 . One of these paths is the one
that is the actual physical path. How does one determine this physical path
(the correct functional form for qi (t ) )?

This is where we use Hamilton’s Principle. Out of all these infinitely possible
paths connecting P1 and P2 , the physical path ( qi (t ) ) would be the one for
102 which the action has the least possible value. In other words the physical
Unit 4 Hamilton’s Principle
path minimizes the action. (It would be more accurate to say that the physical
path makes the action an extremum or stationary). This is called the Principle
of Least Action or “Hamilton’s Principle”, because it was first proposed by
Hamilton in 1834:
t2
S    Lqi (t ), q i (t ), t dt  0 (4.2)
t1

Hamilton’s principle can be stated as follows(Classical Dynamics of Particles


and Systems, Stephen T. Thornton and Jerry B. Marion):

Of all the possible paths along which a dynamical system


may move from one point to another in configuration space
in a given time interval from t1 to t 2 , the actual path along
which the system moves, is that which minimizes the time
integral of the Lagrangian (the action) for the system along
the path.
The time integral of the Lagrangian along the path is also called the action
or action integral for the system. This principle describes the motion of all
mechanical systems for which all forces except the constraint forces are
derivable from some generalized scalar potential function that may be a
function of coordinates, velocities and time.
The following ideas are subsumed in the statement of Hamilton’s principle:
1. All the physical information about a system can be summarized in terms of
a Lagrangian.
2. The paths taken by any system will be paths of stationary action.
Let us now derive the E L equations from Hamilton’s Principle.

4.2.3 Derivation of the Euler-Lagrange Equations of


Motion from Hamilton’s Principle
Let us assume, to start with, that the system has just one degree of freedom
and a single generalized coordinate q. The motion of the system is between
 
the points P1 q 1  and P2 q 2 between the time t  t1 and t  t 2 . So the action
is written as an integral of the Lagrangian function L  Lq, q, t  as:
t2
S   Lq, q, t dt (4.3)
t1

Let us now say that q  q(t ) is the function that minimizes S. q  q(t )
therefore defines the actual path between the points P1 and P2 in
configuration space. Now let us say that q(t ) is replaced by a function
q(t )  q(t ) , where q(t ) is a function which is small over the entire interval of
time from t  t1 to t  t 2 , and is zero at t  t1 and t  t 2 . So now we have a
path between P1 and P2 which is different from the physical path through an
arbitrary(though small) variation of q  q(t ) at each point which is represented
by q(t ) . The paths coincide at the end points of the path so the condition on
this new path is : 103
Block 1 The Lagrangian Formulation of Mechanics
q(t1 )  q(t 2 )  0 (4.4)

We can write the corresponding change in the action as:


S  S q  q, q  q   S q, q 
t2 t2


 Lq  q, q  q, t dt  Lq, q, t dt  (4.5)
t1 t1

Now we use the Taylor series expansion in two variables (q and q ) for
Lq  q, q  q, t  to determine how the Largrangian changes for an
infinitesimal change in the path, i.e., when q changes by q :

L L 
Lq  q, q  q, t   Lq, q, t   q  q
q q
 ... higher order terms in q and q (4.6)

In the limit of q and q being infinitesimal, i.e. in the limit q , q  0 but not
exactly equal to zero, we can drop the higher order terms in the expansion of
Lq  q, q  q, t  and write:
t2 t2
 L L  
S    Lq, q, t  
 q
q 
q  
q dt  Lq, q, t dt
t1 t1

t2
 L L 
   q q  q q dt (4.7)
t1

Let us now integrate the second term in the integral of Eq. (4.7) by parts. We
dq  d
get (using    q  and q(t1 )  q(t 2 )  0 ):
 dt  dt
t2 t2 t2
 L    L  dq    L d
  q dt      dt   q dt 
 q   q  dt    q dt 
t1 t1 t1

t t2
L 2
d  L 

q
q 
t1
 dt  q qdt
t1

t2
d  L 
   qdt (4.8)
t1 dt  q 

Substituting the results of Eq. (4.8) in Eq. (4.7) we get:


t2 t2
 L d  L    L d  L 
 q 
S   q   q  dt  
dt  q   
   qdt
q dt  q   (4.9)
t1 t1

The condition for q  q(t ) to be the function the minimizes S ( S  0 ) is


t2
 L d  L 
  q  dt  q qdt  0 (4.10)
t1

104
Unit 4 Hamilton’s Principle
This integral has to be zero for all values of q . For this, the integrand must
be identically equal to zero, and hence the condition of Hamilton’s Principle
reduces to
d  L  L
  0 (4.11)
dt  q  q

which is the Euler Lagrange equation for the generalized coordinate q. This
derivation is for the special case of a system with just one degree of freedom.
For a system with n degrees of freedom,
q i (t1 )  q i (t 2 )  0; i  1,2,3,...n (4.12)
And with each qi being varied independently:
t2 t2

 
S  Lq i  q i , q i  q i , t dt  Lq i , q i , t dt (4.13)
t1 t1

Proceeding in the same way we would get:


t2
 L d  L 
  qi    q i dt  0
dt  q i 
(4.14)
t1

Since all qi ’s are independent of each other we get:

d  L  L
   0, i  1,2,3,...n (4.15)
dt  q i  q i

As you can see, now we have derived the Euler-Lagrange equations of motion
without any reference to Newton’s Laws, directly from Hamilton’s Principle. All
that is required is a knowledge of the Lagrangian function of the physical
system.
The Hamilton’s Principle is therefore a whole new way of formulating
mechanics. It also has wide ranging applications beyond the realm of
mechanics.
Example 4.1
a) For a Lagrangian Lqi , q i , t  defined as

dF
Lqi , q i , t   Lqi , q i , t   ,
dt
where F q i , t  is any arbitrary function of the generalized coordinates
and time, which is differentiable with respect to the generalized
t2
coordinates, show that :   Lqi (t ), q i (t ), t dt  0 .
t1

b) Show that the form of Hamilton’s Principle can be preserved under a


generalized coordinate transformation.
Solution : a) We can write

105
Block 1 The Lagrangian Formulation of Mechanics
t2 t2 t2 t2
dF 
 L  dt    L 
  
 dt   L dt  F (q i , t )
 dt 
t1 t1 t1 t1

F F
where F  q i  t .
q i t
t2
At the end points we know that q i  t  0 and we know that  Ldt  0 . 
t1


Thus we have  L  dt  0 .

b) Let us say that we have transformed from the set time q1, q 2 ,..., q n , t
to Q1,Q2 ,...,Qn ,  . The transformation equations for the generalized
coordinates and time is:
q i  q i (Q1,Q2 ,...,Qn ) (i)

t  t (Q1,Q2 ,...Qn , ) (ii)

Using Eqs. (i) & (ii) in the original expression for the Lagrangian
L  L(q i , t ).
We get a function of Qi and  , L  L (Q1,Q2 ,...,Qn , ).

Now,
dt
  
 L dt   L (Qi , ) dt   L (Qi , )
d
d

dt
Now suppose we write L  L (Qi , )
d
dt
 
We get  L dt   L d and for L  L (Qi , )
d
we can conserve the form

of the E-L equations of motion.

4.2.4 Generalized Momentum


Let us start with the example of the system studied in Sec. 3.4.1 with the
following Lagrangian:
1  2  2 2
L  T V  mx  y  z   V ( x, y , z ) (4.16)
2
You can see that:
L L L
 mx;  my ;  mz (4.17)
x y z

The linear momentum along the x, y and z directions are respectively


p x  mx, p y  my and pz  mz , hence we have

L L L
 px ;  py ;  pz (4.18)
x
 y
 z

106
Unit 4 Hamilton’s Principle
Let us now extend this concept of linear momentum to a system described by
a set of generalized coordinates {q i } , i  1,2,3,...n by defining a generalized
momentum corresponding to each generalized coordinate, as follows:
L
pi  , i  1,2,3,...n (4.19)
q i
Note that because the generalized coordinate q i need not have the
dimension of length (as in the Cartesian coordinate system), the generalized
momentum pi , will not, in general have the dimension of linear momentum
that is familiar to you i.e. kg ms 1. You can see that for yourself in the
following SAQ.

SAQ 2
Determine the generalized momenta for each of these systems:
a) For the simple pendulum of mass m and length L0 which has the following
Lagrangian:
1
L mL0 2  2  mgL0 cos 
2
b) For the spherical pendulum of mass m and length L0 (TQ 3, Unit 3), which
has the Lagrangian:
1
L m L0 2 (  2   2 sin 2 )  mg L0 cos 
2

The generalized momentum is also called the “canonical momentum” or the


“conjugate momentum”.
Further, if you have a velocity dependent potential U  U qi 
, q i , t  , as
defined in Sec. 3.3, then even if qi corresponds to a Cartesian coordinate, pi
will not have the dimension of the linear momentum in Cartesian coordinates.
You can see that in the following example
Example 4.2
Consider the following Lagrangian for the electric charge
 q of mass m moving

in a charge free region containing an electric field E and a magnetic field B :

mx  y  z   q  q Ax x  Ay y  Az z 
1  2  2 2
L
2
Determine the generalized momenta p x , py and pz .

Solution : Using Eq. (4.19) with q1  x, q 2  y and q 3  z, we get


L  1
m x 2  y 2  z 2   q  q Ax x  Ay y  Az z 

px  
x x  2 
 mx  qAx
Similarly
L
py   my  qAy
y

107
Block 1 The Lagrangian Formulation of Mechanics
L
And pz   mz  qAz .
z

Next we define a function analogous to the energy that you are familiar with.

4.2.5 Energy Function


Let us start by determining the rate of change of the Lagrangian function of the
system L  Lqi , q i , t  :

dL L  dq i  L  dq i  L
dt
      
q i  dt  i q i  dt  t
(4.20)
i

From the Euler-Lagrange equations we get


L d  L 
   (4.21)
q i dt  q i 
L
Substituting for from Eq. (4.21) into Eq. (4.20) we can write:
q i

dL d  L  L  dq i  L
dt
  
dt  q i
q i 

  
q i  dt  t
(4.22)
i i

Now because:
d  L    d  L  L  dq i 
 qi     q i    (4.23)
dt  q i   dt  q i  q i  dt 

We can write Eq. (4.22) as:


dL d  L   L
dt
   qi  
dt  q i  t
(4.24)
i

Which can also be written as:

d   L    L

  qi   L 
dt  i  q i 
0 (4.25)
 t
 L 
The quantity   q i q i   L is called the “energy function” of the system and
i
denoted by h. In general,
 L 
h  hq i ,q i , t     q i q i   L   pi q i   L (4.26)
i i

h can be a function of the generalized coordinates, the generalized velocities


and time. We can therefore rewrite Eq. (4.25) as:
dh L
 (4.27)
dt t
With these definitions, let us now try to see what one can deduce about the
system from the functional form of the Lagrangian.

SAQ 3

108 Write down the energy function for the following systems:
Unit 4 Hamilton’s Principle
a) A bead of mass m moving on an elliptical wire (SAQ 3a, Unit 3) with the
Lagrangian:
1
L  T  V  m(a 2 sin 2  b 2 cos2 ) 2  mgb sin 
2
b) A spherical pendulum(SAQ 2b) the Lagrangian:
1
L m L0 2 (  2   2 sin 2 )  mg L0 cos 
2

4.3 CONSERVATION THEOREMS


4.3.1 Cyclic Coordinates and Conservation of Generalized
Momentum
In general, we know that the Lagrangian is an explicit function of all
generalized coordinates and velocities, and time, that is L  Lqi , q i , t  . Let us
consider a particular case in the Lagrangian is not an explicit function of a
particular generalized coordinate, say q k . Then q k is called a cyclic
coordinate or an ignorable coordinate for the system. Note that the Lagrangian
could still have an explicit dependence on the corresponding generalized
velocity q k .

To understand what a cyclic coordinate tells us about the system we use the
Euler-Lagrange equation for q k :

d  L  L
   0 (4.28)
dt  q k  q k
Because L does not depend explicitly on q k , we can write

L
0 (4.29)
q k

Using Eq. (4.29) in Eq. (4.28) we get:


d  L  L
    0    pk  constant (4.30)
dt q
 k qk

In other words, the generalized momentum pk conjugate to the cyclic


coordinate qk is a constant of motion or pk is conserved.

In Sec. (4.4) you will study how this conservation principle is related to the
underlying symmetries of the system.

Example 4.3

Write down the Lagrangian for projectile motion and determine (i) the cyclic
coordinates and (ii) the conserved momenta.

Solution : Since the motion of the particle is in a plane, we can choose the
plane of motion to be the xy-plane, the generalized coordinates can be taken
as x and y (z = 0). The Lagrangian is:

mx  y   mgy
1 2 2
L
2
109
Block 1 The Lagrangian Formulation of Mechanics
i) The cyclic coordinate is x, since there is no x term in the Lagrangian.

ii) The conserved momentum is


L  1
px    m x 2  y 2   mgy 
x x  2
  

 mx

4.3.2 Conservation of the Energy Function


Notice that, if the Lagrangian has no explicit dependence on time, that is
L
L  Lq i , q i  , then  0 and
t
dh
0 (4.31)
dt
So the energy function will be a constant. In other words, if L is not an explicit
function of time, the energy function h  
 pi q i   L is conserved. This
i
constant is sometimes called the Jacobi integral.
4.3.3 Integrals of Motion
In general, for any mechanical system, any function of qi  and q i  which
remains constant throughout the motion of the system, are called an integral
of motion. Remember that this may happen in spite of the fact that qi  and
q i vary with time. How many such integrals of motion can a system have?
For a system with n degrees of freedom, this number is 2n  1. Let us see
how.
For a system with n degrees of freedom, there are n generalized coordinates,
q1, q 2 ,...., qn and n E-L equations of motion, which are all second order
differential equations. The solution then contains 2n constants of motion
which are determined through the 2n initial conditions, which could be the
initial values of the generalized coordinates and velocities of the system. Let
us say that these 2n constants of motion are c1, c 2 ,...., c 2n . In general we can
write the generalized coordinates and velocities as functions of these 2n
constants and time as follows:
q1  q1t , c1, c 2 ,...., c 2n 

q1  q1t , c1, c 2 ,...., c 2n 

 (4.32)
qn  q n t , c1, c 2 ,...., c 2n 

q n  q n t , c1, c 2 ,...., c 2n 

Now let us say that we pick any one such equation and invert it to obtain the
expression for t in terms of the variable ( qi or q i ) and the 2n constants of
motion. Then we get (for example):
t  t q i , c1, c 2 ,...., c 2n 

110
Unit 4 Hamilton’s Principle
Now we can substitute for t in each of the remaining 2n  1 equations to get
equations containing the generalized coordinate/velocity and constants of
motion. In each of these you could write some function of generalized
coordinates and velocities as a constant,with no explicit dependence on time.
In this way you could generate 2n  1 functions of qi and q i which are
constants of motion.
You can have a maximum of 2n  1 such functions, or 2n  1 integrals of
motion.
With these definitions, let us now try to see what one can deduce about the
system from the functional form of the Lagrangian.

4.4 SYMMETRY- HOMOGENEITY AND ISOTROPY


In the previous section we talked about the conservation laws related to
motion, and you can see that they are similar to the conservation of
momentum and energy you studied about in Unit 1. In Unit 1, you studied that
although these conservation laws are of fundamental importance, they
appeared as an outcome of the dynamical laws that govern nature.
Historically, the notion that conservation laws are a manifestation of the
underlying symmetries that govern these very laws of nature developed much
later, only in the later half on the 20th century. An important consequence of
the existence of symmetries in physics is the existence of laws of
conservation.
In general, if the system or any physical property of the system does not
change under some operation on the system, the system is said to have a
symmetry with respect to that particular operation. Intrinsic to the symmetries
that we define in classical mechanics is the notion of a “closed” system. You
have studied about closed systems in thermodynamics. A closed system is a
physical system that doesn't exchange any matter with its surroundings,
and isn't subject to any net force whose source is external to the system.
We now show how each conservation law is related to a fundamental
symmetry of the system. We talk about three important symmetries:
 Homogeneity of Time: If the physical properties of the system remain
unchanged for an arbitrary displacement of the origin of time, then we say
that time is homogeneous.
 Homogeneity of Space: If the physical properties of the system remain
unchanged for any arbitrary displacement of the origin of reference frame ,
then we say that space is homogeneous.
 Isotropy of Space: If the physical properties of the system remain
unchanged for any arbitrary rotation of the reference frame about its origin,
then we say that space is isotropic.
We already know that Newton’s law are valid only in inertial frames of
reference, in which space is both homogeneous and isotropic, and time is
homogenous (Unit 1, Sec. 1.2). Therefore, the action and hence the
Lagrangian of the system, from which the equations of motion of a system are
derived must also be invariant under rotations and translations in space and
111
Block 1 The Lagrangian Formulation of Mechanics
time translation. Let us see how these symmetries of the Lagrangian lead us
to the familiar conservation principles.
a) Homogeneity of time and energy conservation
If the Lagrangian of a dynamical system has to be invariant under time
translation it means that the Lagrangian cannot have an explicit time-
L
dependence. Therefore  0 , L  Lqi , q i  and and we can write
t
dh
 0  h  constant (4.33)
dt
So the energy function does not change with time. In other words, if L is
not an explicit function of time, the energy function (or the Jacobi integral
 L  
as it is also called) : h   q i  L 
q i 

 pi q i   L is conserved. Under
i  i
some special circumstances the energy function also represents the total
energy of the system:
i) You have seen that, general the kinetic energy of the system can be
written as:

T  T0  T j q j  T jk q j q k (4.34)
j j ,k

However, for a system with holonomic and scleronomic constraints, the


transformation equations do not depend explicitly on time and the
kinetic energy is a homogenous quadratic form in the generalized
velocities:

T  T jk q q j k (4.35)
j ,k

ii) Further let us say that the potential V does not depend on the
generalized velocities or time, but just on the generalized coordinates,
that is V  V {q i }  .

With these conditions:

 L   (T  V )     
  q i q i    q i 

q i 
q i
 
 T q q q
jk j k i
i i i  j ,k 
  
   q i V {qi } q i (4.36)
i

Since T jk is a function only of generalized velocities, we can write:

         
 q i  T jk q j q k q i   T jk  q i q j q k  T jk q j  q i q k q i
i  j ,k  i  j ,k j ,k 

 
  
 T jk  ij q k  T jk q j ik q i
i  j ,k j ,k 

112
Unit 4 Hamilton’s Principle

 
  Tik q k  T ji q j q i
i k j 
 2T (4.37)
V
Since V  V {qi }  ,  0. So the energy function can be written as:
q i
h  2T  L  2T  (L  V )  T  V (4.38)

This is just the total mechanical energy of the system. So the


homogeneity of time leads us to the law of conservation of
energy.
b) Homogeneity of space and conservation of linear momentum

The homogeneity of space implies that the mechanical properties of a


closed system cannot change if the entire system undergoes a parallel
displacement. Let us now consider a translation of the system as a whole
in a given direction, so that the generalized coordinate qi changes to
qi  dq i . You can think of this as a shift in the origin of the coordinate
system. Since the velocities in the system should not change due to a shift
in the origin, the kinetic energy of the system T should also not change
with a shift in the origin of the coordinate system. So we can assume that T
does not depend on qi . We also assume that V is a function only of the
generalized coordinates and not of the velocities. Then in Eq. (3.26) of
T
Unit 3 we can write, with 0:
qi

d  T 
   Qi (4.39)
dt  q i 

where Qi is generalized force along the ith generalized coordinate:

3N
 x 
Qi   Fk A  qki  (4.40)
k 1

This equation can also be written as:


N  
A  rk 
Qi  
Fk .  
 qi 
(4.41)
k 1

Now (see Fig. 4.1) because of the translation of the vector rk changes to

rk and we can write:
  
rk rk (qi  dq i )  rk (q i )
 Lim  nˆ (4.42)
q i dq i  0 dq i

where n̂ is the unit vector along the direction of the translation of the
system. Therefore Eq. (4.41) reduces to:
N 
Qi   Fk A . nˆ (4.43)
k 1
113
Block 1 The Lagrangian Formulation of Mechanics
Therefore Qi is the net applied force along the direction of translation of
the system. Further

1 N 1 N  
T  
2 k 1
mk rk 2  
2 k 1
mk rk .rk (4.44)

z
And the conjugate momentum for the ith generalized coordinate is :

dqi nˆ  
N   rk  N
T    mk v k . rk 

pi    mk rk .
 q 
 q i k 1  q i   i 
rk (qi )   k 1
N  N 

rk  (qi  dqi )
  mk v k .nˆ  nˆ.  mk v k (4.45)
k 1 k 1
y
which is the component of the total linear momentum of the system
x along the direction of translation. So finally we have from Eqs. (4.41)
and (4.45):
Fig. 4.1: Translation.
d
 pi   Qi (4.46)
dt
Where the rate of change of the total linear momentum of the system in
any direction is equal to the total applied force in that direction. Now if V is
V
not a function of qi ( qi is a cyclic coordinate) , we get Qi    0 and
q i
Eq. (4.46) reduces to
d
 pi   0 (4.47)
dt
Which is just the familiar statement of the law of conservation of linear
momentum, which says that if the net applied force in any direction is zero
then the total linear momentum of the system in that direction is zero.
To summarise, if the generalized coordinate representing the translation of
the system is a cyclic coordinate, it would indicate that the translation of
the system has no effect on the dynamics of the system. So if your system
remains invariant under a particular translation, the total linear momentum
of the system along the direction of translation is conserved.
In a similar fashion, we can show that if dqi corresponds to a rotation of
the system as a whole about some axis, then if qi is a cyclic coordinate,
the corresponding angular momentum of the system is conserved.
c) Isotropy of space and conservation of angular momentum
The isotropy of space implies that the mechanical properties of a closed
system of particles cannot change if the entire system is rotated as a
whole. We assume as in (b) that T is a function only of the generalized
velocities and V is a function only of the generalized coordinates. As
before, let us now consider that this rotation of the system is represented
by a change in the generalized coordinate from qi to qi  dq i , except that
now qi corresponds to an angular variable. So, the generalized force is
N  
 r 
114
Qi   Fk A .  k 
 q i 
(4.48)
k 1
Unit 4 Hamilton’s Principle
Let us say that under an infinitesimal rotation dq i (see Fig. 4.2), the vector
 
rk changes to rk with the magnitude of the vector remaining constant, so
z
we get: dq i
  
drk   rk   rk  rk sin dq i

(4.49)

Thus 
 rk (qi )
rk 
q i
 rk sin  (4.50)
 rk  (qi  dqi )
 y
r
The direction of k is at a tangent to the path shown by the dotted line
qi
x
and hence perpendicular to both the unit normal n̂ and the position vector

rk , we represent this direction by the unit vector tˆ . The cross product of n̂ Fig. 4.2: Rotation.

and rk is:
 
nˆ  rk  nˆ rk sin  tˆ  rk sin  tˆ (4.51)
So 
rk 
 nˆ  rk (4.52)
q i

Using the results of Eq. (4.52) in Eq. (4.48) and using the properties of the
scalar triple product, we can write
N 

k 1
 N  
Qi   Fk A . nˆ  rk    nˆ. rk  Fk A
k 1
  (4.53)

 
Notice that rk  Fk A is just the torque due to force acting on the kth particle
about the origin of the coordinate system. Let us denote this torque by

k . Then Eq. (4.53) is:
N N
 
Qi   
nˆ.  k  nˆ.  k (4.54)
k 1 k 1
This understanding of
So Qi is the component of the total applied torque about the axis of the connection between
symmetry and
rotation. (The generalized force need not necessarily be a force, just as the
conservation laws went
generalized coordinate need not be a length). Once again, unnoticed until 1918,
  when Emmy Noether
T N   rk  N   r k 
  mk rk .
 q i   k k  q i 
 m v .  proved her famous
q i k 1   k 1 theorem relating
symmetry and
N   N  
  
mk v k .nˆ  rk   nˆ. (rk  mk v k ) conservation laws. You
can read about the
k 1 k 1
importance of symmetry
N   in:
 nˆ.  rk  pk  (4.55)
David J. Gross, Proc.
k 1
  Natl. Acad. Sci, USA,
Identifying rk  pk as the angular momentum of the kth particle about the Vol 93, pp. 14256-14259,
 T December 1996.
origin, which we can denote by Lk , is just the component of the total
q i
angular momentum along the axis of rotation. So the Euler-Lagrange
equation in this case reduces to:
115
Block 1 The Lagrangian Formulation of Mechanics
d
 Li    i (4.56)
dt
Now if qi is a cyclic coordinate, i  0 and the angular momentum about
the axis of rotation is a constant of motion.
If the generalized coordinate representing the rotation of the system is a
cyclic coordinate, it would indicate that the rotation of the system does not
change the dynamics of the system and therefore the conjugate
generalized momentum along the axis of rotation is conserved. In this case
the conjugate momentum is the angular momentum of the system.
Therefore we have recovered the laws of conservation of linear and
angular momentum from the underlying symmetry of the system.
We will study symmetry in more detail in the next semester, where we talk
about Noether’s Theorem.

4.5 HAMILTON’S PRINCIPLE FOR


NONHOLONOMIC SYSTEMS
We have emphasized that the E-L equations of motion hold for systems with
holonomic constraints. In an N-particle system with no constraints 3N
independent coordinates are required to specify the position of the system at
any instant. If the system has k holonomic constraints, then if
3N  k coordinates are known to us the remaining k coordinates can be
determined using the k constraint relations. Therefore the system has 3N  k
degrees of freedom and 3N  k independent generalized coordinates.
This is not true in a system with nonholonomic constraints. Suppose a system
has k holonomic constraints and k1 nonholonomic constraints, then the
number of degrees of freedom in the system is n  3N  k  k1 but the number
of generalized coordinates required is n   3N  k  n  k1 . So the number of
generalized coordinates are greater than the number of degrees of freedom in
the system ( n  n ) and these generalized coordinates are now not all
linearly independent. The number of independent generalized coordinates
is equal to the number of degrees of freedom which is n. Thus we cannot
derive E-L equations, assuming the linear independence of all generalized
coordinates (Sec. 3.2.3).
We now show how it is still possible to derive Hamilton’s principle for a special
class of nonholonomic constraints using the method of Lagrange
undetermined multipliers. Let us consider that the nonholonomic constraints in
the system are defined by k1 equations of the kind:
f p  f p (q1, q 2 ,....q n  , q1, q 2 ,....q n  )  0, p  1,2,....k1 (4.57)

Eq. (4.57) also implies that


k1
  pfp  0 (4.58)
p 1

where  i are undetermined quantities, which can be functions of the


generalized coordinates and time. If we assume that Hamilton’s principle holds
116 for this system we also have:
Unit 4 Hamilton’s Principle
t2
  Lqi (t ), q i (t ), t dt  0 (4.59)
t1
And
t2
 L d  L 
  q i  
dt  q i
 q i dt  0

(4.60)
t1

where now all qi are not independent on account of the nonholomic
L d  L 
constraints and therefore Eq. (4.60) does not imply     0.
q i dt  q i 
However, combining Eqs. (4.58) and (4.59) we can still write:
t2  k1 
  Lqi (t ), q i (t ), t     p fp  dt  0 (4.61)
t1  p 1 

Let us now carry out the variation over n   n  k1 variables which are the n
independent generalized coordinates q1, q2 ,....qn and k1 undetermined
constants  p .

So now, the modified Hamilton’s principle reads:


t2
 F q1, q 2 ,....q n , 1, 1,.... k1 , t dt  0
 (4.62)
t1

k1
where F  Lqi (t ), q i (t ), t     pf p (4.63)
p 1

Using Eq. (A3.2) of Appendix 3A we can write:


  k1   k1 
 
 
 pfp   

 pfp 
 
d  L  d   p 1  L  p 1  0
     (4.64)
dt  q i  dt 
 q i
  q i q i
 
 

Using the further simplification that  p   p (t ) , we get

d  L  d p   ( f p )  d  f p 
  
dt  q i  p dt  q i  p
 
 p  
dt  q i 

L ( f p )

qi
 
p
qi
0 (4.65)
p

Which is
d  L  L  ( f p ) d  p  ( f p )  d  f p 

dt  q i

 q i
   p q i
 
dt  q i 
   p 
dt  q i


p

(4.66)
Writing

117
Block 1 The Lagrangian Formulation of Mechanics
  (f p )
d  p  ( f p )  d  f p 
Qi  
 p q  dt  q    p dt  q  (4.67)
p  i  i   i 

Using Eq. (4.67) in Eq. (4.66)) we can write


d  L  L
   Qi (4.68)
dt  q i  q i
Where Qi is the generalized force. If the variation over  p is carried out we can
also determine these undetermined multipliers. These help us to determine
the forces of constraint and you shall see how in Example 4.4.
Let us now work out an example on this method of variational principle applied
to a system with nonholonomic constraints.

Example 4.4
Determine the equations of motion for the following Lagrangian of a particle
moving in a plane (the generalized coordinates are x and y):
1
L m( x 2  y 2 )  V ( x, y )
2
Subject to the nonholonomic constraint f ( x, y , x, y )  xy   x  0 .

Solution : Using Eq. (4.66) we can write the following two equations for x
and y:
d  L  L f  f d f
        (i)
dt  x  x x x
 dt  x 
d  L  L f  f d  f 
and        (ii)
dt  y  y y x dt  y 

where  is the Lagrange multiplier for the single nonholonomic constraint


defined by f ( x, y , x, y ) .

 From Eqs. (i) & (ii) we get for the given Lagrangian and f :
V
mx      y   y (iii)
x
and
V
my    x   y (iv)
y

The other equation is:


 x y   x  0 (v)

Eqs. (iii), (iv) and (v) represent the three equations for the three unknowns
x, y ,  .

In this Unit we have derived the laws of motion for a dynamical system using a
variational principle, the Hamilton’s Principle. We have said that the true path
of motion is one that optimizes a quantity, called the action. This is your
introduction to the notion that a law of nature can be formulated from a
118 variational principle and the workings of the physical world are in some sense
Unit 4 Hamilton’s Principle
optimal, which is a whole new idea. This is not how you have been stating
physical laws.
Recollect that you derive the dynamics of a system from Newton’s laws, you
have Maxwell’s laws for electric and magnetic fields, Einstein’s equations for
gravitational field and the equations for the fields of elementary particles. Yet
the principle of least action seems to be a more fundamental principle
because this single equation leads us to the classical equations of motion for
particles and as you will study later, all four of Maxwell’s equations and plays a
central role in quantum mechanics (you can read Feymann Lectures Vol. II,
Chapter 19 for an overview on the ramifications of this important principle).
In the context of the dynamics of a system from the principle of least action,
remember that this principle deals only with conservative systems. There
are no dissipative forces like friction included in this formalism and energy is
conserved. This is a statement of the fundamental truth that on a microscopic
level, energy is truly conserved.
With this Unit we complete our study of the Lagrangian formalism. In the
remaining part of the course you will study two important applications of this
formalism, namely central forces and small oscillations.

4.6 SUMMARY
 Configuration Space
For an N-particle dynamical system with n degrees of freedom we can
define an n-dimensional configuration space in which each point is
labelled by the set of n generalized coordinates, q1, q 2 ,.......q n . As the
dynamical system evolves with time, it traces out a path in
configuration space.
 Action
The action S ( sometimes referred to as the action integral) for any
physical system along any path traversed by the physical system
between two points in configuration space, is a functional which is the
line integral of the Lagrangian function along the path:
t2
S   Lqi (t ), q i (t ),t dt
t1

 Hamilton’s Principle

Of all the possible paths along which a dynamical system may move
from one point to another in configuration space in a given time interval
from t1 to t 2 , the actual path along which the system moves, is that
which minimizes the time integral of the Lagrangian( the action) for the
system along the path:
t2
S    Lqi (t ), q i (t ), t dt  0
t1

The Euler-Lagrange Equations of Motion


119
Block 1 The Lagrangian Formulation of Mechanics
d  L  L
   0, i  1,2,3,...n
dt  q i  q i

can be derived from Hamilton’s Principle.

 Generalized Momenta and Energy Function


For a dynamical system described by a set of generalized coordinates
{qi } , i  1,2,3,...n we define a generalized momentum( also called a
conjugate momentum or canonical momentum) corresponding to
each generalized coordinate, as follows:
L
pi  , i  1,2,3,...n
q i

The energy function (h) of the dynamical system is:

h  hq i ,q i , t    pi q i   L


i

 Cyclic Coordinates and Conservation of Generalized Momentum


If the Lagrangian is not an explicit function of a particular generalized
coordinate, say q k , q k is called a cyclic coordinate for the system
and the generalized momentum pk conjugate to the cyclic
coordinate q k is a constant of motion for the system.

 Conservation of the Energy Function


If the Lagrangian has no explicit dependence on time, the energy
dh
function will be a constant: 0
dt

 Symmetry: Homogeneity and Isotropy


The homogeneity of time leads to energy conservation for a system
with conservative forces.
The homogeneity of space leads to the conservation of linear
momentum.
The isotropy of space leads to the conservation of angular momentum.
 Hamilton’s Principle for Nonholonomic System
For a system with nonholonomic constraints defined by k1 equations:

f p  f p (q1, q2 ,....qn  , q1, q 2 ,....q n  )  0, p  1,2,....k1

Hamilton’s Principle reduces to


t2  k1 
  Lqi (t ),q i (t ),t     pf p  dt  0
t1  p 1 

Where  p are undetermined quantities (Lagrange multipliers), which


can be functions of the generalized coordinates and time. The Euler-
Lagrange equations of motion have the form:

120
Unit 4 Hamilton’s Principle

d  L  L  ( f p ) d  p  ( f p )  d  f p 

dt  q i

 q i
   p q i
 
dt  q i 
   p 
dt  q i


p

4.7 TERMINAL QUESTIONS


1. Determine the generalized momenta and energy for
a) Bead of mass m sliding on a long straight were rotating with an
constant angular speed  at a constant angle 0 with the vertical,
which has the following Lagrangian (Unit 3, TQ 5):
1
L  m(r 2   2 r 2 sin 2  0 )  mgr cos  0 )
2
b) Block of mass m sliding down a moving frictionless inclined plane of
mass M (Unit 3, SAQ 3b) which has the Lagrangian
1 1
L  T V  M  m  X 2   ms 2  mX s cos   mg s sin   mga.
2 2

c) A pendulum with of mass m and length L0 with a point of support


moving vertically. The motion of the support is described by the
function d(t).
2. The Lagrangian for a particle of mass m moving freely on the surface of a
sphere of radius A has the following Lagrangian
1
L m A 2 (  2  sin 2   2 )
2
Write down the equations of motion. Determine the generalized momenta
p and p and the cyclic coordinates. Show that the Lagrangian and p
are constants of motion
3. A disc of radius a is rolling without slipping down a frictionless inclined
plane which makes an angle  with the horizontal. Write down the
Lagrangian and the nonholonomic equation of constraint. Obtain the
equations of motion.
4. A mass m is suspended from a point on the celling using a spring of spring
constant k and unstretched length L0. Write down the Lagrangian and
obtain the generalized momenta and the energy integral.

4.8 SOLUTIONS AND ANSWERS


Self-Assessment Questions
1. a) The generalized coordinates are q1   and q 2   . So the
configuration space is two dimensional with 0    180 and
0    360 .

b) There is just one generalized coordinate q1  y1 . So the configuration


space is one-dimensional and the range of the coordinate is 0  y1  l .

2. a) For the generalized coordinate , using Eq. (4.19)


L  1 2 2 
p 


  mL0   mgL0 cos  
   2  121
Block 1 The Lagrangian Formulation of Mechanics
2
 mL0 

b) For the generalized coordinates  and , using Eq. (4.19) we get


L  1 
p   m L0 2 ( 2   2 sin2 )  mg L0 cos 
   2 
 mL0 2

L  1 
p    m L0 2 ( 2   2 sin 2 )  mg L0 cos 
 
   2 
 mL0 2 sin2 

3. a) Using Eq. (4.19) we first calculate the generalized momentum


L   1 
p   m(a 2 sin 2   b 2 cos 2 ) 2  mgb sin 
   2 

 m(a 2 sin 2   b 2 cos 2 )

We determine the energy function using Eq. (4.26)


h  p   L  m(a 2 sin 2   b 2 cos 2 ) 2
1
 m(a 2 sin 2  b 2 cos 2 ) 2  mgb sin 
2
1
 m(a 2 sin 2  b 2 cos 2 ) 2  mgb sin 
2
b) Using the results of SAQ 2b and Eq. (4.26), we can write:
h  p   p  L

1
 m L0 2  2  m L0 2  2 sin 2   m L0 2 ( 2   2 sin 2 )  mg L0 cos 
2
1
 mL0 2 ( 2   2 sin 2 )  mg L0 cos 
2

Terminal Questions
1. a) Using Eq.(4.19) we get:
L  1 
pr    m(r 2   2 r 2 sin 2  0 )  mgr cos  0   mr
r
 r  2
 

Using Eq. (4.26) we get


1 2 1
h  p r r  L  mr  m  2 r 2 sin 2  0  mg r cos  0
2 2
b) Using Eq. (4.19) we get
L  1  2 1 2 
 2 M  m X  2 ms  mX s cos   mg s sin   mga 
pX    
 
X X  

 M  m  X  m s cos 

L  1
ps    M  m X 2  1 ms 2  mX s cos   mg s sin   mga 
122
s s  2
  2 
Unit 4 Hamilton’s Principle

 ms  m X cos 

The energy function is (Eq. 4.26)


h  p X X  ps s  L
  
 MX  mX  m s cos  X  ms  m X cos  s 
1 1 
  M  m X 2  ms 2  mX s cos   mg s sin   mga 
 2 2 
1 1
 M  m X 2  ms 2  mX s cos   mg s sin   mga
2 2

c) See Fig. 4.3. We can write the coordinates of the bob as d (t )


x
x  L0 sin   x  L0 cos ; y  L0 cos   d (t )  y  L0 sin   d O
L0

L
1 2 2
2
 
m x  y  mg (L0 cos   d )
y

1 x
 m(L0 2 2  2L0d sin   d 2 )  mg (L0 cos   d )
2
y
The generalized momentum is (Eq. 4.19)
Fig. 4.3
L
p   mL0 2  m L0 d sin 


The energy function (Eq. 4.26) is
h  p   L
1 
 mL0 2  2  mL0 d sin    m(L0 2  2  2L0 d sin   d 2 )  mg (L0 cos   d ) 
 2 
1
 m(L0 2  2  d 2 )  mg (L0 cos   d )
2
2. The equations of motion are
d  L  L
   0  mA 2  mA 2 sin  cos  2  0 (i)
dt    

And
d  L  L
    0  2mA 2 sin  cos     mA 2 sin2    0 (ii)

dt    
The generalized momenta are
L L
p   mA2 ; p    mA2 sin2 

 

The cyclic coordinate is  . So the generalized momentum p  mA 2 sin 2   is


a constant of motion. For L to be a constant of motion, we must have
dL L
 0 . Since  0 we have
dt t
dL L  L  L  L 
     
dt    
123
Block 1 The Lagrangian Formulation of Mechanics

 mA 2 sin  cos    2  mA 2    mA 2 sin2    (iii)

Using mA 2  mA 2 sin  cos  2 from Eq. (i) and


mA 2 sin 2    2mA 2 sin  cos    from Eq. (ii) in Eq. (iii) we get

dL
 mA 2 sin  cos    2  mA 2 sin  cos   2  2mA 2 sin  cos    2  0
dt
 L is a constant of motion

3.The generalized coordinates are x and  as shown in Fig. 4.4.


The nonholonomic constraint is ad  dx from which we write the following
equation of constraint:
f ( x, )  x  a   0 (i)

The Lagrangian is:


y
1 2 1
L mx  ma 2 2  mg h1
2 2
a
x
h1  (h  x sin  )
 1 2 1
 L  mx  ma 2 2  mg (h  x sin  ) (ii)
2 2
h1
 Since we have one nonholonomic constraint, we introduce one Lagrange
x multiplier  to get the following equation
Fig. 4.4: TQ 3.
d  L  L f  f d  f 
         mx  mg sin   
dt  dx  x x x dt  x 

Or mx  mg sin     0 (iii)

And
d  L  L f  f d  f 
         ma 2  a (iv)
dt  d     dt   
Eqs. (i), (iii) and (iv) are three equations for the unknowns r,  and 

From Eq. (i) we have x  a  x  a


 x 
 ma 2  a  ma 2    a  or   mx (v)
O x a
r
 y
Substituting from Eq. (v) in Eq, (iii) we get
x g sin 
mx  mg sin   mx  0  x  (vi)
2

y and   g sin  (vii)


2a
Fig. 4.5: TQ 4
4. The generalized coordinates are r and  (Fig. 4.5).
Using x  r sin  and y  r cos 
124
Unit 4 Hamilton’s Principle
We get
x  r cos    r sin  ; y  r sin    r cos 

 The kinetic energy is: T 


1
2
1

m( x 2  y 2 )  m r 2   2 r 2
2

1 1
The potential energy is: V  mgy  k (r  L0 )2  mgr cos   k (r  L0 ) 2
2 2
1 2 1
The Lagrangian is: L  m(r  r 2 2 )  mgr cos   k (r  L0 )2
2 2
The generalized momenta are
L L
pr   mr ; p   mr 2 
r 
The energy function is:
1 1
h  pr r  p  L  m(r 2  r 2  2 )  mgr cos   k (r  L0 ) 2
2 2

125
Block 2 Applications of Quantum Mechanics

FURTHER READINGS

1. Classical Mechanics, N. C. Rana, P.S. Joag, Tata McGraw –Hill


Publishing Company Limited (1992).

2. Classical Mechanics, Herbert Goldstein, Charles Poole, John Safko, 3 rd


Edition, Pearson Education (2002).

3. Mechanics, L. D. Landau and E. M. Lifshitz, 3rd Edition, Pergamon


Press (1976).

4. Classical Mechanics, R. Douglas Gregory, Cambridge University Press


(2006).

126

You might also like