0% found this document useful (0 votes)
70 views136 pages

Phy 301

Uploaded by

Pelumi Ajiboye
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
70 views136 pages

Phy 301

Uploaded by

Pelumi Ajiboye
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 136

COURSE

GUIDE

PHY 301
CLASSICAL MECHANICS II

Course Team Mr. A. A. Musari (Course Developer/Writer) –


Crescent University, Abeokuta.
Dr. S. O. Ajibola (Programme Leader –
NOUN
Dr. Efiong A. Ibanga (Course Coordinator) –
NOUN

NATIONAL OPEN UNIVERSITY OF NIGERIA


PHY 301 CLASSICAL MECHANICS II

© 2023 by NOUN Press


National Open University of Nigeria
Headquarters
University Village
Plot 91, Cadastral Zone,
Nnamdi Azikiwe Expressway
Jabi, Abuja

Lagos Office
14/16 Ahmadu Bello Way
Victoria Island, Lagos

e-mail: [email protected]
URL: www.nouedu.net

All rights reserved. No part of this book may be reproduced, in any


form or by any means, without permission in writing from the publisher.

Printed 2023

ISBN: 978-058-005-0

Reviewed 2023
PHY 301 CLASSICAL MECHANICS II

Introduction

PHY 301 Classical mechanics is a first semester course. It is a three(3) credit


course available to all students and learners offering Bachelor of Science
(B.Sc.) Physics.

The course comprises of 11 study units (4 modules) which involve basic


principles of classical mechanics.

Classical mechanics is the study of the motion of bodies (including the special
case in which bodies remain at rest) in accordance with the general principles
first enunciated by Sir Isaac Newton in his Philosophiae Naturalis Principia
Mathematica (1687), commonly known as the Principia. Classical mechanics
was the first branch of Physics to be discovered, and is the foundation upon
which all other branches of Physics are built.

Moreover, classical mechanics has many important applications in other areas


of science, such as Astronomy (e.g., celestial mechanics), Chemistry (e.g., the
dynamics of molecular collisions), Geology (e.g., the propagation of seismic
waves, generated by earthquakes, through the Earth's crust), and Engineering
(e.g., the equilibrium and stability of structures).

Classical mechanics is also of great significance outside the realm of science.


After all, the sequence of events leading to the discovery of classical
mechanics-starting with the ground-breaking work of Copernicus, continuing
with the researches of Galileo, Kepler, and Descartes, and culminating in the
monumental achievements of Newton-involved the complete overthrow of the
Aristotelian picture of the Universe, which had previously prevailed for more
than a millennium, and its replacement by a recognizably modern picture in
which humankind no longer played a privileged role.

This course has a compulsory pre-requisite of PHY 201 So the learners are
strongly advised to have adequate knowledge of analytical mechanics.

This course guide tells you briefly what the course is all about, what course
materials you will be using and how to work your way through these materials.
It suggests some general guidelines for the time to complete it successfully. It
gives you some guidance on your tutor marked assignments.

There are regular tutorial classes that are linked to the course. You are linked to
the course. You are advised to attend these sessions regularly. Details of time
and locations of tutorials will be given to you at the point of registration for the
course.
PHY 301 CLASSICAL MECHANICS II

What You Will Learn In This Course

The overall aim of PHY 301 is to introduce and to explain the concept of
constraints, generalized coordinates, motion under central conservative forces,
scattering, Kepler’s law, Motion in non-inertia frames of reference, the
Lagrange and Hamilton’s formulation of mechanics.
The course consists of units and a course guide. This course guide tells you
briefly what the course is all about, what course materials you will be using and
how you can work with these materials. In addition, it advocates some general
guidelines for the amount of time you are likely to spend on each course in
order to complete it successfully.

Course Aims

The aim is to introduce you to the fundamental principles and concepts of


classical mechanics and their application in everyday life. This will be achieved
by
 Introducing you to the concept of constraints and generalized coordinates.
 Explaining the motion of two bodies interacting via a central conservative
force.
 Explaining the importance applications of Lagrangian mechanics and
conservation theorems.
 Explaining Hamiltonian formulation of mechanics.
 Explaining the concept of motion under non-inertia reference frame.

Course Objectives

The course sets overall objectives, to achieve the aim set out above.

In addition each unit also has specific objectives. The unit objectives are
always included at the beginning of a unit; you should read them before you
start working through the unit. You may want to refer to them during your
study of the unit to check your progress. You should always look at the unit
objectives after completing a unit. In this way you can be sure that you have
done what was required of you for the unit.

Set out below are the objectives of the course a whole. By meeting these
objectives you should have achieved the aims of the course as a whole.

On successful completion of the course, you should be able to:

1. Explain and distinguish between different classes of constraints and give


examples each.
2. Express many Physical quantities in terms of generalized coordinates.
3. Choose appropriate generalized coordinates in any physical system.
PHY 301 CLASSICAL MECHANICS II

4. Calculate the Lagrangian and the corresponding Lagrange’s equation of


motion of many physical system.
5. Calculate the Hamiltonian and the corresponding Hamilton’s equation of
motion of many physical system.
6. Derive the d’Alembert’s principle.
7. Transform Lagrangian in another coordinate system.
8. Describe motion under central conservative force.
9. State Kepler’s Laws and its major application in the case of Gravity
10. Explain Scattering cross section.
11. Determine the time derivatives of motion in fixed and rotating reference
frame.
12. Explain the motion relative to earth and its application in a free falling
object.

Appendix

Working through this material


To complete this course you are required to read the study units, read set
textbooks and read materials provided by NOUN. Each of the units contains
self assessment exercises, and certain points in the course you would be
required to submit assignments herein referred to as Tutor Marked
Assignments (TMAs) for assessment purposes. At the end of the course there
is a final examination. The course should take you about a total of 17 weeks to
complete. Below you will find listed all the components of the course, what
you have to do and how you should allocate your time to each in order to
complete the course and successfully.

This course entails that you spend a lot of time to read. I would advice that you
avail yourself the opportunity of attending the tutorial sessions where you have
the opportunity of comparing your knowledge with that of other learners.

Course Materials

1. Course guide
2. Study units
3. Assignment file
4. Presentation schedule

Study Units

There are 11 Study units in this course as follows:


Module 1 Generalized coordinates and constraints
Unit 1 Constraints
Unit 2 Generalized Coordinates
Unit 3 Virtual work, Virtual Displacement and Generalized Forces
PHY 301 CLASSICAL MECHANICS II

Module 2 Lagrange’s and Hamilton’s formulation of mechanics


Unit 1 d’Alembert’s Principle of Virtual Work
Unit 2 Lagrangian Mechanics
Unit 3 Hamiltonian Mechanics

Module 3 Central force and scattering


Unit 1 The Generic Central Force Problem
Unit 2 Kepler’s Problem
Unit 3 Scattering Cross Section

Module 4 Motion in non-inertia reference frame


Unit 1 Time Derivative in Fixed and Rotating Frames
Unit 2 Motion Relative to Earth

Each study unit consists of two to three weeks work and includes specific
objectives. Each unit contains a number of self-tests. In general, these self-
tests, question you on the material you have just covered or require you to
apply in some ways and thereby, help you to gauge your progress and reinforce
your understanding of te material. Together with tutor marked assignments,
these exercises will assist you in achieving the stated learning objectives of the
individual units and of the course.

Set Textbooks

L.D. Landau and E.M. Lifshitz (1979): Course of Theoretical Physics.


Mechanics Vol. 1 3rd ed. Translated from Russian Press, UK.

K.R. Symon (1974)Mechanics, 3rd ed., Addison-Wesley, USA.

T.W.B. Kibble (1973): Classical Mechanics 2nd ed., McGraw-Hill, UK.

M.R. Spiegel (1982): Schaum’s outline of theory and problems of Theoretical


Mechanics. SI (Metric) Edition with an introduction to Lagrange’s Equations
and Hamiltonian Theory McGraw-Hill, UK.

H.Goldstein (2002) Classical Mechanics, Narosa Publishing Home, New Delhi.

Marion and Thomtron (2000) Classical Dynamics of Particles and Systems,


Third Edition, Horoloma Book Jovanovich College Publisher.

R.G.Takawale and P.S.Puranik, Tata Mc-Graw (1998) Introduction to Classical


Mechanics by Hill Publishing Company Limited, New Delhi
PHY 301 CLASSICAL MECHANICS II

Assignment File

The assignment file will be supplied by NOUN. In this file you will find the
details of the work you must submit to your tutor for marking. The marks you
obtain for these assignments will count towards the final mark you obtain for
this course. Further information on assignments will be found in the assignment
file itself and later in this course guide in the section on assessment.

Presentation Schedule

The presentation schedule included in your course materials may show the
important dates for the completion of tutor marked assignment. Remember, you
are required to submit all your assignments by the due date as dictated by your
facilitator. You should guide against falling behind in your work.

Assessment

There are two aspects to the assessment of the course. First are the tutor-
marked assignments: second, there is a written examination.

In doing the assignment, you are expected to apply information, knowledge and
techniques gathered during the course. The assignments must be submitted to
your tutor for formal assessment in accordance with the dead-lines stated in the
presentation schedule and the assignment file. The work you submit to your
tutor for assessment will count for 30% of your total course work.

At the end of the course you will need to sit for a final written examination of
three hours duration. This examination will also count for 70% of your course
mark.

Tutor-Marked Assignment (TMA)

The TMAs are listed in each unit. Generally, you will be able to complete your
assignments from the information contained in the study units, set books and
other reading. However, it is desirable in all degree level education to
demonstrate that you have read and researched more widely than the required
minimum. Using other references will give you a broader viewpoint and may
provide a deeper understanding of the subjects.

When you have completed each assignment, send it, together with a TMA
form, to your tutor. Make sure that each assignment reaches your tutor on or
before the deadline given in the presentation schedule and assignment file. If,
for any reason you cannot complete your work on time contact your tutor
before the assignment is due to discuss the possibility of an extension.
Extensions will not be granted after the due date unless there are exceptional
circumstances.
PHY 301 CLASSICAL MECHANICS II

Final Examination and Grading


The final examination for PHY 301 will be three hours duration and have a
value of 70% of the total course grade. The examination will consist of
quantities which reflect the types of self testing practice exercises and tutor-
marked problems you have previously encountered. All areas of the course will
be assessed.

You are advised to use the time between finishing the last unit and sitting for
the examination to revise your self-tests, tutor-marked assignments and
comments on them before the examination.

Course Marking Scheme

The table shows how the actual course marking is broken down.

Assessment Marks
Assignments 30% of course marks
Final Examination 70% of overall course marks
Total 100% of course marks

Table 1: course marking scheme

Facilitators/Tutors and Tutorials

There are 16 hours of tutorials provided in support of this course. You will be
notified of the dates, times and location of these tutorials as well as the name
and phone number of your facilitator, as soon as you are allocated a tutorial
group

Your facilitator will mark and comment on your assignments, keep a close
watch on your progress and any difficulties you might face and provide
assistance to you during the course. You are expected to mail your Tutor
Marked Assignment to your facilitator before the schedule date (at least two
working days are required). They will be marked by your tutor and returned to
you as soon as possible.

Do not delay to contact your facilitator by telephone or e-mail if you need


assistance.

The following might be circumstances in which you would find assistance


necessary, hence to contact your facilitator if:
 You do not understand any part of the study or the assigned readings.
 You have difficulty with self-tests.
 You have a question or problem with an assignment or with the grading
of an assignment.
PHY 301 CLASSICAL MECHANICS II

You should endeavour to attend the tutorials. This is the only chance to have
face to face contact with your course facilitator and to ask questions which are
answered instantly. You can raise any problem encountered in the course of
your study.

To gain much benefit from course tutorials prepare a question list before
attending them. You will learn a lot from participating actively in discussions.

Summary

Classical mechanics is a course that intends to provide the concept of motion of


bodies in accordance with general principles firs enunciated by Sir Isaac
Newton. It is the most common system of physics today. It is the physics of
“ordinary” situations, considering objects too large to exhibit quantum effects,
too slow to exhibit relativistic effects and not too dense enough to require
general relativity. Upon completing this course, you will be equipped with the
basic understanding of constraints, generalized coordinates, Lagrange’s and
Hamilton’s formulation of mechanics and motion in non-inertia frame of
reference. In addition, you will be able to answer the following type of
questions.
 Define degree of freedom and constraint.
 Identify and Explain the three kinds of non-holonomic constraints.
 Identify and Explain the two kinds of holonomic constraint.
 Define Degree of freedom
 Define constraints.
 Mention two categories of Constraints and Give three examples each.
 Mention three kinds of non-holonomic constraints.
 Show that velocity dependent constraints are non-integrable constraints.
 Write Velocity of any system with cartesian coordinates x and y in
generalized coordinates.
 Express Kinetic energy of a particle with Cartesians coordinates x and y
in a generalized coordinates.
 Derive generalized force from the expressions of virtual work and
virtual displacement.
 State and Derive the D’Alembert’s Principle of virtual work.
 
Given that kinetic energy T   mi ri  ri  T q k , q k , t  derive the
1

i 2
generalized equation of motion.
 Given U ( z)  mgz where z  b sin calculate the generalized force.
 From the generalized equation of motion prove the Euler-Lagrangian
equation and so on

Of course, the list of questions you can answer is not limited to the above list.
To gain the most from this course you should endeavour to apply the
PHY 301 CLASSICAL MECHANICS II

Lagrange’s and Hamilton’s formulation of mechanics to every other system


like Atwood’s machine, Elliptical wire, central force system, motion of the
earth, free fall systems and so on.

I wish you success in the course and I hope that you will find it both interesting
and useful.
MAIN
COURSE

CONTENTS

Module 1 Generalized Coordinates and Constraints ……… 1

Unit 1 Constraints ………………………………………… 1


Unit 2 Generalized Coordinates…………………………... 9
Unit 3 Virtual work, Virtual Displacement
and Generalized Forces……………………………. 14

Module 2 Lagrange’s and Hamilton’s


Formulation of Mechanics………………………. 23

Unit 1 D’Alembert’s Principle of Virtual Work……………. 23


Unit 2 Lagrangian Mechanics……………………………. 33
Unit 3 Hamiltonian Mechanics…………………………… 53

Module 3 Central force and Scattering …………………… 65

Unit 1 The Generic Central Force Problem …………………. 65


Unit 2 Kepler’s Problem ………………………………… 82
Unit 3 Scattering Cross Section …………………………. 96

Module 4 Motion in Non-Inertial Reference Frame ……… 107

Unit 1 Time Derivative in Fixed and Rotating Frames ……. 107


Unit 2 Motion Relative to Earth………………………….. 115
PHY 301 CLASSICAL MECHANICS II

MODULE 1 GENERALIZED COORDINATES AND


CONSTRAINTS

Unit 1 Constraints
Unit 2 Generalized Coordinates
Unit 3 Virtual work, Virtual Displacement and Generalized Forces

UNIT 1 CONSTRAINTS

CONTENTS

1.0 Introduction
2.0 Objectives
3.0 Main Contents
3.1 Degrees of Freedom
3.2 Constraints
3.2.2 Non-holonomic Constraints
3.2.1 Holonomic Constraints
4.0 Conclusion
5.0 Summary
6.0 Tutor Marked Assignment
7.0 References/Further Reading

1.0 Introduction

A rigid body is defined as a system of n particles for which all the


 
interparticle distances are constrained to fixed constants, ri  r j  cij and
the interparticle potentials are functions only of these interparticle
distances. As these distances do not vary, neither does the internal
potential energy. These interparticle forces cannot do work, and the
internal potential energy may be ignored.

The rigid body is an example of a constrained system, in which the


general 3n degrees of freedom are restricted by some forces of constraint

which place conditions on the coordinates ri perhaps in conjunction with
their momenta. In such descriptions we do not wish to consider or
specify the forces themselves, but only their (approximate) effect. It is
generally assumed, as in the case with the rigid body, that the constraint
forces do no work under displacements allowed by the constraints

1
PHY 301 CLASSICAL MECHANICS II

2.0 Objectives

By the end of this unit, you will be able to:


 Define degree of freedom and constraint.
 Distinguish between holonomic and non-holonomic constraints.
 Give examples of holonomic and non-holonomic constraints.
 Identify and Explain the three kinds of non-holonomic constraints.
 Identify and Explain the two kinds of holonomic constraints

3.0 Main Contents

3.1 Degrees of Freedom

The number of degrees of freedom is defined as the number of


independent coordinates that is needed to identify uniquely the
configuration of the system.
Suppose we have a system of N particles each moving in 3-space and
interacting through arbitrary (finite) forces, The number n=3N is called
the number of degree of freedom.
There is freedom, of course, in how we specify the degrees of freedom;
e.g.:
 choice of origin
 coordinate system: cartesian, cylindrical, spherical, etc.
 center-of-mass vs. individual particles: rk  or R, sk  rk  rk  R
    

But, the number of degrees of freedom is always the same; e.g., in the

center-of-mass system, the constraint k sk  0 applies, ensuring that
 
rk  and R , sk have same number of degrees of freedom.
The motion of such a system is completely specified by knowing the
dependence of the available degrees of freedom on time.
Simple illustration are: ―The beads of a abacus are constrained to one-
dimensional motion by the supporting wires‖ so also ―Gas molecules
within a container are constrained by the walls of the vessel to move
only inside the container‖.
Throughout this section, we will work two examples alongside the
theory. The first consists of an Atwood’s machine, and we may allow
the rope length to be a function of time, l = l(t)
The second consists of point particle sliding on an elliptical wire in the
presence of gravity. The Cartesian coordinates of the particle satisfy
2 2
 x   z 
      1
 a(t )   b(t ) 
(1.0)
We will at various points consider a and b to be time dependent or
constant. The origin of the coordinate system is the stationary centre of
the ellipse.

2
PHY 301 CLASSICAL MECHANICS II

Figure 1.1 (a) Atwood machine (b) Point particle sliding


on elliptical wire

Example 1.1:
In the Atwood’s machine of figure 1.1(a), there are a priori 2 degrees of
freedom; the z coordinates of the two blocks. (We ignore the x and y
degrees of freedom because the problem is inherently 1-dimensional.)
The inextensible rope connecting the two masses reduces this to one
degree of freedom because, when one mass moves by a certain amount,
the other one must move by the opposite amount.

Example 1.2:
In the elliptical wire of figure 1.1(b), there are a priori 3 degrees of
freedom, the 3 spatial coordinates of the point particle. The constraints
reduce this to one degree of freedom, as no motion in y is allowed and
the motions in z and x are related. The loss of the y degree of freedom is
easily accounted for in our Cartesian coordinate system; effectively, a
2D Cartesian system in x and z will suffice. But the relation between x
and z is a constraint that cannot be trivially accommodated by dropping
another Cartesian coordinate.
Self Assessment Exercise A

3
PHY 301 CLASSICAL MECHANICS II

1. Now define Degree of freedom


2. State the number of degrees of freedom in an Atwood’s machine
and elliptical wire

3.2 Constraints

Constraints mean a restriction on the degree of freedom of motion of a


system of particles in the form of a condition.
Constraints may reduce the number of degrees of freedom; e.g., particle
moving on a table, rigid body, etc. It is divided into Holonomic and
Non-Holonomic constraint.

3.2.1 Holonomic Constraints

Holonomic constraints are those that can be expressed in the form


 
f (r1 , r2 ,t )  0 . For example, restricting a point particle to move on the
surface of a table is the holonomic constraint z  z0  0 , where z0 is a
constant. A rigid body satisfies the holonomic set of constraints
 
ri  r j  cij  0 , where cij is a set of constants satisfying cij  c ji  0 for
all particle pairs i, j.
A system is called ―holonomic‖ if, in a certain sense, one can recover
global information from local information, so the meaning ―entire-law‖
is quite appropriate.
Holonomic constraints may be divided into rheonomic (―running law‖)
and scleronomic
(―rigid law‖) depending on whether time appears explicitly in the
constraints:

rheonomic: f rk , t   0 ,
(1.1a)

Scleronomic: f rk   0 .
(1.1b)
f
At a technical level, the difference is whether  0 or not: the presence
t
of this partial derivative affects some of the relations we will derive later.

3.2.2 Non-holonomic Constraints

The rolling of a ball on a table is non-holonomic, because one rolling


along different paths to the same point can put it into different
orientations.
Non-holonomic constraints are, obviously, constraints that are not
holonomic. There are three kinds of non-holonomic constraints:
i Non-integrable or history-dependent constraints. These are
constraints that are not fully defined until the full solution of the
equations of motion is known. Equivalently, they are certain

4
PHY 301 CLASSICAL MECHANICS II

types of constraints involving velocities. The classic case of this


type is a vertical disk rolling on a horizontal plane. If x and y
define the position of the point of contact between the disk and
the plane, ϕ defines the angle of rotation of the disk about its
axis, and ϴ defines the angle between the rotation axis of the
disk and the x-axis, then one can find the constraints
x   r cos ,
y  r sin  .
(1.2a)
The differential version of these constraints is
dx   rd cos ,
dy   rd sin  .
(1.2b)
These differential equations are not integrable; one cannot generate from
the relations two equations f1 x, ,    0 and f 2  y, ,    0 . The reason
is that, if one assumes the functions f1 and f2 exist, the above differential
equations imply that their second derivatives would have to satisfy
2 f 2 f
 i.e for f  f1or f 2
 
(1.3)
which is very unpleasant mathematical condition. Explicitly, suppose f1
existed. Then we would be able to write f1 ( x, , )  0
Let us obtain the differential version of the constraint by differentiation:
f1 f f
dx  1 d  1 d  0 .
x  
(1.4)
This differential constraint should match the original differential
constraint dx  rd cos . Identifying the coefficients of the differential
yields
f1 f1 f1
1 0  r cos .
x  
(1.5)
Taking the mixed second partial derivatives gives
 2 f1  2 f1
0  r sin
 
(1.6)
which, clearly, do not match. Such constraints are also called non-
integrable because one cannot integrate the differential equation to find
a constraint on the coordinates. A differential relation such as the one
above is a local one; if the differential relation is integrable, you can
obtain the constraint at all points in space, i.e., you can find the ―entire
law‖. Clearly, non-integrability is also related to the fact that the
constraint is velocity-dependent: a velocity-dependent constraint is a

5
PHY 301 CLASSICAL MECHANICS II

local constraint, and it may not always be possible to determine a global


constraint from it.
ii. inequality constraints; e.g., particles required to stay inside a
box, particle sitting on a sphere but allowed to roll off,
iii. problems involving frictional forces.

Example 1.3:
For the elliptical wire example in figure 1.1(b), the constraint equation is
the one we specified initially:
2 2
 x   z 
      1
 a(t )   b(t ) 
If a and/or b do indeed have time dependence, then the constraint is
rheonomic. Otherwise, it is scleronomic.

Example 1.4:
For the Atwood’s machine in figure 1.1(a), the constraint equation is
z1  z2  l (t )  0
Where l(t) is the length of the rope (we assume the pulley has zero
radius). The signs on z1 and z2 are due to the choice of direction for
positive z in the example. The constraint is again rheonomic if l is
indeed time dependent, scleronomic if not

Self Assessment Exercise B


1. Define constraints.
2. Mention two categories of Constraints and Give three examples
each.
3. Mention three kinds of non-holonomic constraints.
4. Show that velocity dependent constraints are non-integrable
constraints.

4.0 Conclusion

Holonomic constraints are constraints that can be written in the form


  
f r1 , r2 rM , t   0 ,
i.e., there is some sort of condition on the coordinates and possibly time.
The condition may not
involve the coordinate velocities. Holonomic constraints are also termed
integrable because a differential version of the constraint can be
integrated to yield the full constraint. The constraint is rheonomic if time
appears explicitly, scleronomic if not.

6
PHY 301 CLASSICAL MECHANICS II

5.0 Summary

 The number of degrees of freedom is defined as the number of


independent coordinates that is needed to identify uniquely the
configuration of the system.
 Constraints means a restriction on the degree of freedom of motion
of a system of particles in the form of a condition
 Constraints may reduce the number of degrees of freedom; e.g.,
particle moving on a table, rigid body, etc.
 Constraints are divided into holonomic and non-holonomic
constraint
 Holonomic constraints are those that can be expressed in the form
 
f (r1 , r2 ,t )  0
 Holonomic constraints may be divided into rheonomic (―running
law‖) and scleronomic
 Non-holonomic constraints are, obviously, constraints that are not
holonomic.
 There are three kinds of non-holonomic constraints: Non-integrable
or history-dependent constraints, inequality constraints and
problems involving frictional forces

6.0 Tutor Marked Assignment

1. Explain briefly what is meant by degrees of freedom.


2. How many degrees of freedom does a rigid body in the following
three dimensional motion has and explain what is meant by the
motion: i) heaving, ii) surging, iii)swaying, iv) rolling,
v) pitching, vi) yawing.
3. How many degrees of freedom are in atwood machine of figure
1.1(a) and elliptical wire of figure 1.1(b).
4. State the constraint equation of the elliptical wire example in figure
1.1(b), and state if it is scleronomic or rheonomic.
5. In the Atwood’s machine in figure 1.1(a), State the constraint
equation and what condition could make the constraint equation
rheonomic.
6. Show that velocity dependent constraints are non-integrable
constraints.

7.0 Further Reading and Other Resources

Hand, L. N. and Finch, J. D. (1998). Analytical Mechanics. Cambridge


University Press.

Thornton, S.T. and Marion, J. B. (2014.). Classical Dynamics of


Particles and Systems, Fifth Edition. Thomson Books/Cole

7
PHY 301 CLASSICAL MECHANICS II

Goldstein, H., Poole, C., and Safko, J. (2000). Classical


Mechanics, Third Edition. Horoloma Book Jovanovich College
Publisher.

Wolfgang, N. (2023). Theoretical Physics 2: Analytical Mechanics


https://issuu.com/yakibooki/docs/theoretical_physics_2_analytical_
mechanics
http://nicadd.niu.edu/~jahreda/phys300/phys300%20Chapter%201%20a
nd%20intro.pdf
https://cdn.preterhuman.net/texts/science_and_technology/physics/Mech
anics/Analytical%20mechanics%20-%20Hand,%20Finch.pdf

Video Link:
https://www.youtube.com/playlist?list=PLtUBquogGkRukyAAWOmq2
LmQx8Yf9H_3E

8
PHY 301 CLASSICAL MECHANICS II

UNIT 2 GENERALIZED COORDINATES

CONTENTS

1.0 Introduction
2.0 Objectives
3.0 Main Contents
3.1 Generalized Coordinates
3.2 ―Dot Cancellation‖
4.0 Conclusion
5.0 Summary
6.0 Tutor Marked Assignment
7.0 References/Further Reading

1.0 Introduction

In general, if one has j independent constraint equations for a system of


M particles with 3M degrees of freedom, then the true number of
degrees of freedom is 3M − j. There is dynamical behaviour of the
system in only these remaining degrees of freedom. One immediately
asks the question that since there are fewer degrees of freedom than
position coordinates, is there any way to eliminate those unnecessary
degrees of freedom and thereby simplify analysis of the mechanical
system? In our example, why keeping both x and z if one of them would
suffice? This question leads us to the idea of generalized coordinates,
which are a set of 3M − j coordinates that have already taken into
account the constraints and are independent, thereby reducing the
complexity of the system.

2.0 Objectives

By the end of this unit, you will be able to:

 Write Physical quantities in terms of generalized coordinates.


 Use constraints equation to define different generalized scheme.
 Calculate azimuthal angle.
 Solve Related Problems.

3.0 Main contents

3.1 Generalized Coordinates

For holonomic constraints, the constraint equations ensure that it will


always be possible to define a new set of 3M −j generalized coordinates
{qk} that fully specify the motion of the system subject to the constraints
and that are independent of each other. The independence arises because

9
PHY 301 CLASSICAL MECHANICS II

the number of degrees of freedom must be conserved. The constraint


equations yield (possibly implicit) functions
 
r1  r1 (q1 , q2 ,q3M  j , t )
(2.1)
that transform between the generalized coordinates and the original
coordinates. It may not always be possible to write these functions
analytically. Some of these coordinates may be the same as the original
coordinates, some may not; it simply depends on the structure of the
constraints. We will refer to the original coordinates as position
coordinates to distinguish them from the reduced set of independent
generalized coordinates. Generalized coordinates are more than just a
notational convenience. By incorporating the constraints into the
definition of the generalized coordinates, we obtain two important
simplifications:
1) the constraint forces are eliminated from the problem; and
2) the generalized coordinates are fully independent of each other. We
shall see the application of these simplifications when we discuss virtual
work and generalized forces.
  d 
Just as the velocity corresponding to a coordinate r j is rj  rj , it is
dt
d
possible to define a generalized velocity q k as qk  qk . Note that in all
dt
cases, velocities are defined as total time derivatives of the particular
coordinate. Remember that if you have a function g  g qk , t  then the
d
total time derivative g is evaluated by the chain rule:
dt
g dqk g
g qk , t   
d
 .
dt qk dt t
(2.2)
It is very important to realize that, until a specific solution to the
equation of motion is found, a generalized coordinate and its
corresponding generalized velocity are independent variables. This can
be seen by simply remembering that two initial conditions which are
qk (t  0) and qk (t  0) are required to specify a solution of Newton’s
second law, because it is a second-order differential equation. Higher-
order derivatives are not independent variables because Newton’s
second law relates the higher-order derivatives to qk  and qk  . The
independence of qk  and qk  is a reflection of the structure of
Newton’s second law, not just a mathematical theorem. Unless
otherwise indicated, from here on we will assume all constraints are
holonomic.

10
PHY 301 CLASSICAL MECHANICS II

Example 2.1:
For the elliptical wire, the constraint equation
2 2
 x   z 
      1
 a(t )   b(t ) 
can be used to define different generalized coordinate schemes. Two
obvious ones are
x and z; i.e., let x be the generalized coordinate and drop the z degree of
freedom, or vice versa. Another obvious one would be the azimuthal
angle α
 z a(t ) 
  tan 1  .
 b(t ) x 
(2.3)

The formal definitions ri qk , t  are then
x  a(t ) cos z  b(t ) sin .
(2.4)
Here, we see the possibility of explicit time dependence in the
relationship between the positions x and z and the generalized
coordinate α.

Example 2.2:
For the Atwood’s machine, either z1 or z2 could suffice as the
generalized coordinate. Let’s pick z1, calling it Z to distinguish the
generalized coordinate, we have
z1  Z z 2  l (t )  Z
This case is pretty trivial.

Self Assessment Exercise A

1. Write Velocity of any system with cartesian coordinates x and y in


generalized coordinates
2. Express Kinetic energy of a particle with Cartesians coordinates x
and y in a generalized coordinates

3.2 ―Dot Cancellation‖

For holonomic constraints, there is a very important statement that we


will make much use of later
 
r r
 .
qk qk
(2.5)
Conceptually, what this says is the differential relationship between a
given degree of freedom
and a generalized coordinate is the same as the differential relationship
between the corresponding velocities. This statement relies on having

11
PHY 301 CLASSICAL MECHANICS II

holonomic constraints. Heuristically,one can understand the rule as


simply arising from the fact that holonomic constraints use only the
positions and time to define the generalized coordinates; as a result, any
relationships between positional velocities and generalized velocities
must be determined only by relationships between positions and
generalized coordinates. The above relationship is then the simplest
possible one.
We derive the result rigorously by starting with the total time

derivative ri :
 
 d  ri dqk ri
r  ri qk , t     .
dt k qk dt t
(2.6)
The last term does not exist if the constraints are scleronomic, but that is
not really important here. Now, take the partial derivative with respect
to q1 ; this selects out the term in the sum with k = l, and drops the t term:
  
r ri ri
  kl  .
ql k qk ql
(2.7)
So the dot cancellation relation is proven.

Self Assessment Exercise B


1. What is Dot cancellation.
 
r r
2. Prove that  .
qk qk

4.0 Conclusion

From the Physicist point of view the q’s are generalized coordinates in
the sense that they need not have dimensions of length. By deriving
equations of motion in terms of a general set of generalized coordinates,
the results found will be valid for any coordinate system that is
ultimately specified.

5.0 Summary

 Given a set of holonomic constraints, one can use the constraints to


implicitly define a set of independent generalized coordinates.
The transformation relations between the generalized coordinates
and the original physical coordinates are of the form
ri  rj q1 , q2 q3M  j , t 
 

where we assume there are j constraint equations. It is assumed that


all the constraints are used so that the generalized coordinates
truly form an independent set.

12
PHY 301 CLASSICAL MECHANICS II

 Dot Cancellation
 
ri ri
For holonomic constraints, it holds that  .
qk qk
6.0 Tutor Marked Assignment

1. Two particles are connected by a rigid rod so they are constrained to


move a fixed distance apart. Write down a constraint equation of the
 
form f (r1 , r2 ,t )  0 and find suitable generalized coordinates for
the system incorporating this holonomic constraint.
Answer
Let the position of particle 1 with respect to a stationary Cartesian frame
be {x1, y1, z1} and that of particle 2 be {x2, y2, z2}. The rigid rod
constraint equation is
x1  x2 2   y1  y2 2  z1  z2 2  z1  z2 2  l 2 .
This reduces the number of degrees of freedom from 6 to 5, which we
could take to be position of particle 1, {x1, y1, z1} However, the three
position coordinates would be the coordinates of the centre of mass
whose position vector is
m2 r1  m1r2
rcm  .
m1  m2

7.0 Further Reading and Other Resources

Hand, L. N. and Finch, J. D. (1998). Analytical Mechanics. Cambridge


University Press.

Thornton, S.T. and Marion, J. B. (2014.). Classical Dynamics of


Particles and Systems, Fifth Edition. Thomson Books/Cole

Goldstein, H., Poole, C., and Safko, J. (2000). Classical Mechanics,


Third Edition. Horoloma Book Jovanovich College Publisher.

Video Link:
https://www.youtube.com/playlist?list=PLtUBquogGkRukyAAWOmq2
LmQx8Yf9H_3E

13
PHY 301 CLASSICAL MECHANICS II

UNIT 3 VIRTUAL DISPLACEMENT, VIRTUAL WORK AND


GENERALIZED FORCES

CONTENTS

1.0 Introduction
2.0 Objectives
3.0 Main Contents
3.1 Virtual Displacement
3.2 Virtual Work
3.3 Generalized Force
4.0 Conclusion
5.0 Summary
6.0 Tutor Marked Assignment
7.0 References/Further Reading

1.0 Introduction

Our discussion of generalized coordinates essentially was an effort to


make use of the constraints to eliminate the degrees of freedom in our
system that have no dynamics. Similarly, the constraint forces, once they
have been taken account of by transforming to the generalized
coordinates, would seem to be irrelevant. We will show how they can be
eliminated in favour of generalized forces that contain only the non-
constraint forces.

To define generalized forces, we combine the equation relating virtual


displacements of position coordinates and generalized coordinates, with
the equation relating virtual work and non-constraint forces as it will be
shown later in this unit.

2.0 Objectives

By the end of this unit, you will be able to:


 Explain virtual displacement, virtual work and generalized forces.

 Express total differential of any set of system position vectors, ri
that are functions of other variables, q1 , q2 ,, qm  and time t in
terms of virtual displacement.
 Solve related Problems.

14
PHY 301 CLASSICAL MECHANICS II

3.0 Main Contents

3.1 Virtual Displacement



We define a virtual displacement ri  as an infinitesimal displacement

of the system coordinates ri  that satisfies the following criteria.
1. The displacement satisfies the constraint equations, but may make
use of any remaining unconstrained degrees of freedom.
2. The time is held fixed during the displacement.
3. The generalized velocities q k  are held fixed during the
displacement.
A virtual displacement can be represented in terms of position
coordinates or generalized coordinates. The advantage of generalized
coordinates, of course, is that they automatically respect the constraints.
An arbitrary set of displacements qk can qualify as a virtual
displacement if conditions (2) and (3) are additionally applied, but an

arbitrary set of displacements ri  may or may not qualify as a virtual
displacement depending on whether the displacements obey the
constraints. All three conditions will become clearer in the examples.
Explicitly, the relationship between infinitesimal displacements of
generalized coordinates and virtual displacements of the position
coordinates is

 ri
ri   q k .
k q k

(3.1)

This expression has content: there are fewer qk  than ri  , so the fact

that ri can be expressed only in terms of the qk  reflects the fact that
the virtual displacement respects the constraints. One can put in any
values of the qk and obtain a virtual displacement, but not every

possible set of ri  can be written in the above way.

Example 3.1:
For the elliptical wire, it is easy to see what kinds of displacements
satisfy the first two requirements. Changes in x and z are related by the
constraint equation; we obtain the relation by applying the
displacements to the constraint equation. We do this by writing the
constraint equation with and without the displacements and
differentiating the two:
2 2
 x  x   z  z 
With displacements       1 ,
 a (t )   b(t ) 
(3.2)

15
PHY 301 CLASSICAL MECHANICS II

2 2
 x   z 
without displacement       1 ,
 a (t )   b (t ) 
(3.3)
2 2 2 2
 x  x   x   z  z   z 
difference             0 ,
 a (t )   a (t )   b(t )   b(t ) 
(3.4)
xx zz x a(t ) z
 0  .
a(t ) 2
b(t ) 2
z b(t ) x
(3.5)
All terms of second order in x or z are dropped because the
displacements are infinitesimal. The result is that x and z cannot be
arbitrary with respect to each other and are related by where the particle
is in x and z and the current values of a and b; this clearly satisfies the
first requirement. We have satisfied the second requirement, keeping
time fixed, by treating a and b as constant: there has been no t applied,
which would have added derivatives of a and b to the expressions. If a
and b were truly constant, then the second requirement would be
irrelevant. The
third requirement is not really relevant here because the generalized
velocities do not enter the constraints in this holonomic case except if
kinetic energy enters the constraints .
The relationship between the virtual displacements in the positions and
in the generalized coordinate is easy to calculate:
x  a cos  x  a sin  ,
(3.6a)
z  b sin   z  b cos ,
(3.6b)

x a
  tan  .

z b
(3.7)
We see that there is a one-to-one correspondence between all
infinitesimal displacements  of the generalized coordinate and virtual
displacements of the positional coordinates x, z  , as stated above. The
displacements of the positional coordinates that cannot be generated
from  by the above expressions are those that do not satisfy the
constraints and are disallowed.

Example 3.2:
For our Atwood’s machine example, the constraint equation
z1  z 2  l (t )  0 is easily converted to differential form, giving
z1  z 2  0
Again, remember that we do not let time vary, so l(t) contributes nothing
to the differential. The equation is what we would have arrived at if we

16
PHY 301 CLASSICAL MECHANICS II

had started with an infinitesimal displacement z of the generalized


coordinate Z
z1  Z z 2  Z  z1  z 2  0 .
Fixing time prevents appearance of derivatives of l(t). Again, also see
the one-to-one correspondence between infinitesimal displacements of
the generalized coordinates and virtual displacements of the position
coordinates.

Self Assessment Exercise A


1. Read and understand the two examples given above properly

3.2 Virtual Work

Using the virtual displacement, we define virtual work as the work that
would be done on the system by the forces acting on the system as the

system undergoes the virtual displacement ri 
 
W   Fij  ri ,
ij

(3.8)
 
where Fij is the jth force acting on the coordinate of the ith particle ri .

Example 3.3:

In our elliptical wire example, F11 would be the gravitational force

acting on the point mass and F12 would be the force exerted by the wire
to keep the point mass on the wire (i = 1 only because there is only one
object involved). The virtual work is
 
1   
W   Fi1  Fi 2  ri .
i 1

(3.9)

Example 3.4:
In our Atwood’s machine example, the two masses feel gravitational
 
forces F11  m1 gzˆ and F21  m2 gzˆ . The tension in the rope is the force
that enforces the constraint that the length of rope between the two
 
blocks is fixed, F21  Tzˆ and F22  Tzˆ . T may be a function of time if l
varies with time, but it is certainly the same at the two ends of the rope
at any instant.

Because of the possibility that there exist situations of the third type, we
use the most generic assumption to proceed with our derivation, which
is the third one. We write
 (c) 
 ij  rij  0 ,
F
ij

(3.10)

17
PHY 301 CLASSICAL MECHANICS II

where the (c) superscript restricts the sum to constraint forces but the sum
is over all constraint forces and all particles. Mathematically, the
assumption lets us drop the part of the virtual work sum containing
constraint forces, leaving
 
W   Fij( nc )  ri ,
ij

(3.11)
where the (nc) superscript indicates that the sum is only over non-
constraint forces.

Example 3.5:

In our elliptical wire example, the force exerted by the wire, F12 , acts to
keep the point mass i on the wire; the force is therefore always normal to

the wire. The virtual displacement r1 must always be tangential to the
 
wire to satisfy the constraint. So, i 1 Fi 2  ri  0 , the only non-constraint
1


force is gravity, F11 , so we are left with
1 1   1    
W    Fij( nc )  ri   Fi1  ri  F11  r1  mgz .
j 1 i 1 i 1

(3.12)
W will in general not vanish; it gives rise to the dynamics of the
problem.

Example 3.6:
 
In the Atwood’s machine example the constraint forces F21 and F22 act
along the rope. The virtual displacements are also along the rope.
   
Clearly, F21  r1  f 21z1 and F22  r2  f 22z 21 do not vanish but the sum
does
2  
F i2  ri  f 21z1  f 22z 2  T z1  z 2   0 .
i 1

(3.13)

Notice that, in this case all the terms pertaining to the particular
constraint force have to be summed in order for the result to hold. The
virtual work and non constraint force sum is
2 1   2  
W   Fij( nc )  ri   Fi1  ri   g (m1z1  m2z 2 )
i 1 j 1 i 1

(3.14)

Self Assessment Exercise B



1. Express total differential of any set of system position vectors, ri
that are functions of other variables, q1 , q2 ,, qm  and time t in
terms of virtual differential displacement

18
PHY 301 CLASSICAL MECHANICS II

ri m
r
dri  dt   i dq j
t j 1 q j

3.3 Generalized Force

To define generalized forces, we combine Equation 3.1, the relationship


between virtual displacements of position coordinates and generalized
coordinates, with Equation 3.11, the relationship between virtual work
and non-constraint forces:
 
W   Fij( nc )  ri
ij

 ( nc )  ri 
  Fij   q k 
ij  k q k 

  r 
   Fij( nc )  i q k
k  ij q k 
  f k q k ,
k

(3.15)
where
 ( nc ) ri W
f k   Fij   ,
ij q k q k
is the generalized force along the kth generalized coordinate. The last
expression, f k  W q says that the force is simple the ratio of the
k

work done to the displacement when a virtual displacement of only the


kth generalized coordinate is performed; it is of course possible to
displace only the kth generalized coordinate because the generalized
coordinates are mutually independent. It is important to remember that
the generalized force is found by summing only over the non-constraint
forces: the constraint forces have already been taken into account in
defining generalized coordinates.
 
Infinitesimally, a force F causing a displacement r does work
 
W  F  r . The generalized force is the exact analogue: if work W is
done when the ensemble of forces act to produce a generalized
coordinate displacement qk , then the generalized force fk doing that
work is f k  W q . But the generalized force is a simplification
k

because it is only composed of the non-constraint forces.

Example 3.7:
In the elliptical wire example, the generalized force for the α coordinate
(k=1) is

19
PHY 301 CLASSICAL MECHANICS II

 r  x z 
f   F11  1  mgzˆ   xˆ  zˆ   mgzˆ   xˆa sin   zˆb cos 
    
 mgb cos .
(3.16)
The constraints force, which acts in both the x and z direction and is α-
dependent, does not appear in the generalized force.

Example 3.8:
In the Atwood’s machine example, the generalized force for the Z
coordinate (k=1 again) is
 r1  r2 z z
f z  F11   F21   m1 g 1  m2 g 2
Z Z Z Z
 m2  m1 g

(3.17)

Again, the constraint force (the rope tension) is eliminated. Because Z is


just z1 in this case, the generalized force in Z is just the net force on m1
acting in the z1 direction.

Self Assessment Exercise C

1. Make sure you understand the above examples very well.


2. Derive generalized force from the expressions of virtual work and
virtual displacement.

4.0 Conclusion

Generalized forces are defined via coordinate transformation of applied


forces, Fi on a system of n particles, i. The concept finds its use in
Lagrangian mechanics, where it plays a conjugate role to generalized
coordinates.
A convenient equation from which to derive the expression for
generalized forces is that of the virtual work, W caused by applied
forces, as seen in D'Alembert's principle in accelerating systems and the
principle of virtual work for applied forces in static systems.
n
W   Fi  ri
i 1

(3.18)

ri is the virtual displacement of the system. Substituting the definition


for the virtual displacement (differential), we have
m
ri
ri   q j ,
j 1 q j

(3.19)

20
PHY 301 CLASSICAL MECHANICS II

n m
ri
then we have W   Fi   q j ,
i 1 j 1 q j

(3.20)
m n
ri
therefore W   Fi  q j .
j 1 i 1 q j
(3.21)
From this form, we can see that the generalized applied forces are then
defined by
W n
r
fj    Fi  i ,
q j i 1 q j
thus, the virtual work due to the applied forces is
n
W   f j q j .
j 1

5.0 Summary

Virtual Displacement and Work



 A virtual displacement is a set of coordinate displacements ri 
that obey the following conditions: 1. The displacement
satisfies the constraint equations
2. Time is held fixed during the displacement.
3. The generalized velocities q k  are held fixed
during the displacement.
 Any displacement of a truly independent set of generalized
coordinates qk  satisfies (1) automatically.
 Virtual work is the work done during a virtual displacement. It is
defined as
 
W   Fij  ri .
ij

Because the displacement is virtual, we make the assumption that


the contribution of constraint forces to the virtual work vanishes,
leaving
 
W   Fij( nc )  ri
ij
(nc)
where refers to non constraint forces only.
 The generalized force along the generalized coordinate qk is
defined to be
 ( nc ) ri
f k   Fi  ,
q k

where Fi (nc ) is the sum of all non-constraint forces acting on
particle i.

21
PHY 301 CLASSICAL MECHANICS II

6.0 Tutor Marked Assignment (TMAs)

1. Derive generalized force from the expressions of virtual work and


virtual displacement

7.0 Further Reading and Other Resources

Hand, L. N. and Finch, J. D. (1998). Analytical Mechanics. Cambridge


University Press.
Thornton, S.T. and Marion, J. B. (2014.). Classical Dynamics of
Particles and Systems, Fifth Edition. Thomson Books/Cole
Goldstein, H., Poole, C., and Safko, J. (2000). Classical Mechanics,
Third Edition. Horoloma Book Jovanovich College Publisher.

Video Link:
https://www.youtube.com/playlist?list=PLtUBquogGkRukyAAWOmq2
LmQx8Yf9H_3E

22
PHY 301 CLASSICAL MECHANICS II

MODULE 2 LAGRANGES AND HAMILTON’S


FORMULATION OF MECHANICS

Unit 1 D’Alembert’s Principle Of Virtual Work


Unit 2 Lagrangian Mechanics
Unit 3 Hamiltonian Mechanics

UNIT 1 D’ALEMBERT’S PRINCIPLE OF VIRTUAL WORK

CONTENTS

1.0 Introduction
2.0 Objectives
3.0 Main contents
3.1 d’Alembert’s Principle
3.2 Generalized Equations of Motion
3.3 The Lagrangian and the Lagrange’s Equations
3.3.1 Generalized Conservative Forces
3.3.2 The Euler-Lagrange Equations
4.0 Conclusion
5.0 Summary
6.0 Tutor Marked Assignment (TMAs)
7.0 Further Reading and Other Resources

1.0 Introduction

D'Alembert's principle is a statement of the fundamental classical laws


of motion. It is named after its discoverer, the French physicist and
mathematician Jean le Rond D’Alembert. The principle states that the
sum of the differences between the forces acting on a system and the
time derivatives of the momenta of the system itself along a virtual
displacement consistent with the constraints of the system, is zero and it
is the basis of Lagrangian Mechanics.
It is the dynamic analogue to the principle of virtual work for applied
forces in a static system and in fact is more general than Hamilton's
principle, avoiding restriction to holonomic systems.
2.0 Objectives
By the end of this unit, you will be able:
 Derive D’Alembert’s Principle from Newton’s Second Law Of
Motion.
 To reformulate Newton’s equations system as a system of
equations for the generalized coordinates q(t).
 Use D’Alembert’s principle to relate generalized forces to the rate
of change of the momenta:
 Solve Related Problems.

23
PHY 301 CLASSICAL MECHANICS II

3.0 Main Contents

3.1 D’Alembert’s Principle

D’Alembert’s Principle: Our definition of virtual work


 
was W   Fij  ri where the sum includes all (constraint and non-

constraint) forces. Assuming our position coordinates are in an inertial


frame (but not necessarily our generalized coordinates), Newton’s

second law tells us ij Fij  pi the sum of all the forces acting on a
particle give the rate of change of its momentum. We may then
rewrite W :
   
W   Fij  ri   p i  ri .
i j i

But, we found earlier that we could write the virtual work as a sum over
 
only non-constraint forces, W   Fij( nc )  ri .

Thus, we may derive the relation


  ( nc )  
  F ij  p   ri  0 ,
i  j 
(1.1)
the above equation 1.1 is referred to as D’Alembert’s principle. The
expression shows that the rate of change of momentum is determined
only by the non-constraint forces. In this form, it is not much use, but
the conclusion that the rate of change of momentum is determined only
by non-constraint forces is an important physical statement.

We may use D’Alembert’s principle to relate generalized forces to the


rate of change of the momenta:

   ri
k Fkqk  W  k p  ri  
i,k
p
qk
qk .

(1.2)

Now unlike the ri , the qk  are mutually independent. Therefore, we
may conclude that equality holds for each of the sum separately
providing a different version of D’Alembert’s principle
 ( nc ) ri 
 ri
ij Fij  q  f k  i p  q .
k k

(1.3)
This is now a very important statement: the generalized force for the kth
generalized coordinate,
which can be calculated from the non-constraint forces only, is related to
a particular weighted sum of momentum time derivatives (the weights
being the partial derivatives of the position coordinates with respect to

24
PHY 301 CLASSICAL MECHANICS II

the generalized coordinates). Effectively, we have an analogue of


Newton’s second law, but including only the non-constraint forces.
This can be a major simplification in cases where the constraint forces
are complicated or simply not known. D’Alembert’s principle is not yet
useful in the current form because the left side contains generalized
forces depending on the generalized coordinates but the right side has
derivatives of the momenta associated with the position coordinates. We
need time derivatives with respect to the generalized coordinates on the
right hand side so that all the dynamics can be calculated in the
generalized coordinates.

Self Assessment Exercise A


1. State and Derive the D’Alembert’s Principle of virtual work

3.2 Generalized Equations of Motion

Here we perform the manipulation needed to make D’Alembert’s


principle useful. We know from Newtonian mechanics that work is
related to kinetic energy, so it is natural to expect
1  
T   mi ri  ri  T q k , q k , t  .
i 2

(1.4)
T should be obtained by first writing T in terms of position velocities
ri  and then using the definition of the position coordinates in terms of
generalized coordinates to re-write T as a function of the generalized
coordinates and velocities. T may depend on all the generalized

coordinates and velocities and on time because the ri  depend on the
generalized coordinates and time and a time derivative is being taken,
which may introduce dependence on the generalized velocities. The
partial derivatives of T are:
 
T  ri  ri
  mi ri    pi  ,
qk i qk i qk
(1.5)
 
T  r  r
  mi ri  i   pi  i ,
qk i qk i qk
(1.6)
where in the last step we have made use of dot cancellation because the

constraints are assumed to be holonomic. Now, we have pi floating

around but we need p i . The natural thing is then to take a time
derivative. We do this to T qk (instead of T qk ) because we want to
avoid second order time derivatives if we are to obtain any expression
algebraically similar to the right hand side of d’Alembert’s principle.
We find

25
PHY 301 CLASSICAL MECHANICS II

 
d  T   ri  d ri
    pi    pi  .
dt  qk  i qk i dt qk
(1.7)
Referring back to the second form of d’Alembert’s principle (Equation
1.1), we see that the first term in the expression is the generalized force
Fk for the kth coordinate. Continuing onward, we need to evaluate the
second term. We have
   
d ri   2 ri  2 ri   2 ri
  ql  ql   .
dt qk l  ql qk ql qk  tqk
(1.8)
When we exchange the order of the derivatives in the second term, we
see that the second term vanishes because our holonomic constraint
assumption that the generalized velocities do not enter the constraints,
and thus do not enter the relationship between position and generalized

coordinates implies ri qk  0. In the last term, we can trivially
exchange the order of the partial derivatives. We can bring  qk outside
the sum in the first term because ql qk  0 . Thus, we have
    
d ri    ri  ri   ri
   ql     ,
dt qk qk 
 l  q l  t 
 q k

(1.9)
where the last step simply used the chain rule for evaluation of
 
ri  dri dt. Essentially, we have demonstrated that the total time
derivative d/dt and the partial derivative  qk commute when acting on

ri for holonomic constraints, which is a nontrivial statement because qk
is time-dependent. We emphasize that it was the assumption of
holonomic constraints that allow us discard the second term above. Had
that term remained, the dependence on q1 would have made it
impossible to bring  qk outside the sum because ql in general may
depend on ql (via Newton’s second law). So we now have

 
d  T   ri  ri
    pi    pi  ,
dt  q k  i q k i q k

 ri T
  pi   ,
i q k q k
(1.10)
or

 ri d  T  T
i pi  q  dt  q   .
k  k  qk
(1.11)
Recalling the d’Alembert’s principle (equation 1.1), we may re-write the
above:

26
PHY 301 CLASSICAL MECHANICS II

 ( nc ) ri d  T  T
i Fi  q  f k  dt  q   .
k  k  q k
(1.12)
This is the generalized equation of motion. The left hand side is
completely determined by the non-constraint forces and the constraint
equations. The right hand side is just derivatives of the kinetic energy
with respect to the generalized coordinates and velocities. Thus, we
obtain a differential equation for the motion in the generalized
coordinates.

Example 1.1:
For the elliptical wire example, the kinetic energy in terms of position
coordinate velocities is
T
2
x  z 2 .
m 2

(1.13)
We have previously obtained formulae for x and z in terms of  :
x  a cos  a sin z  b sin  b cos
(1.14)
Let us specialize to the case a  0 and b  0 ; so that
x  a sin z  b cos
(1.15)
We use these to rewrite the kinetic energy in terms of  :
T
2

m 2 2 2
a  sin   b 2 2 cos2  
(1.16)
This is an important example of how to convert T from a function of the
position velocities to a function of the generalized coordinates and
velocities. Now take the prescribed derivatives:
d  T  T
  
d
dt     dt
    
m a 2 sin 2   b 2 cos2   2m 2 a 2  b 2 sin  cos

 ma sin   b
2 2 2
cos2  
(1.17)
In taking the total derivative in the first term, we obtain two terms: the
one displayed in the last line above, and one that exactly cancels the last
term in the first line. We have the generalized force fα from equation
3.16, fα= −mg b cos α, so the generalized equation of motion is
d  T  T
f   
dt    
(1.18)
Substituting equation 1.17 into equation 1.18, we have
 mgb cos  ma2 sin2   b2 cos2  .
(1.19)

27
PHY 301 CLASSICAL MECHANICS II

b cos
Solving for  , we get    g .
a sin   b 2 cos2 
2 2

(1.20)
Specializing to a = b = r (circular wire), this simplifies to
cos
   g .
r
(1.21)
Example 1.1:
For the Atwood’s machine example, things are significantly simpler.
The kinetic energy is
T
1
2
m1 z12  m2 z22  .
(1.22)
Re-writing using the generalized coordinate Z gives
T
1
m1  m2 Z 2 .
2
(1.23)
The kinetic energy derivatives term is
d  T  T
   m1  m2 Z .
dt  Z  Z
(1.24)
Using fz = (m2 – m1)g from equation 3.17 above, the generalized
equation of motion is
d  T  T
fz   
dt  Z  Z
m2  m1 g  m1  m2 Z
m  m2
Z   1 g.
m1  m2
(1.25)

Self Assessment Exercise B


1. Make sure you study and understand the examples given above
properly.
 
2. Given that kinetic energy T   mi ri  ri  T q k , q k , t  derive the
1
i 2
generalized equation of motion.

3.3 The Lagrangian and the Lagrange’s Equations

For conservative non-constraint forces, we can obtain a slightly more


compact form of the generalized equation of motion, known as the
Euler-Lagrange equations.

28
PHY 301 CLASSICAL MECHANICS II

3.3.1 Generalized Conservative Forces

Now let us specialize to non-constraint forces that are conservative; i.e.


 
Fi ( nc )  iU rj 

 
where  i indicates the gradient with respect to ri . Whether the
constraint forces are conservative is irrelevant; we will only explicitly
need the potential for the non-constraint forces. U is assumed to be a
function of the coordinate positions only; there is no explicit dependence

on time or on velocities, U t  0 and U ri  0 . Let us use this
expression in writing out the generalized force:
 ( nc ) ri  
ri 
   iU r j 

f k   Fi   U ql , t 
i q k i q k q k
. (1.26)
We made use of the holonomic constraints to re-write U as a function of
the {ql} and possibly t, and realize that the previous line is just the
partial derivative of U with respect to qk. Thus, rather than determining
the equation of motion by calculating the generalized force from the
non-constraint forces and the coordinate transformation relations, we
can re-write the potential energy as a function of the generalized
coordinates and calculate the generalized force by gradients thereof.

Example 1.2:
For the elliptical wire the potential energy is due to gravity,
U ( z)  mgz .
(1.27)
Re-writing in terms of α, gives
U ( , t )  mgb(t ) sin .
(1.28)
The generalized force is then
U ( ; t )
f   mgb(t ) cos .

(1.29)
as obtained in equation 3.16. Note that we may allow b to be a function
of time without ruining the conservative nature of the potential energy U
becomes a function of t through the definition of the generalized
coordinate but, obviously, if it is initially a conservative potential, a
transformation of coordinates cannot change that.

Example 1.3:
For the atwood’s machine, the potential energy function is
U ( z1 , z2 )  g m1 z1  m2 z2 
(1.30)
Re-writing in terms of Z gives

29
PHY 301 CLASSICAL MECHANICS II

U Z , t   g m1  m2 Z  m2l (t )
(1.31)
The generalized force is
U Z ; t 
fz    g m2  m1 
Z
(1.32)
as found earlier. Again, l is allowed to be function of time without
ruining the conservative nature of the potential energy.

3.3.2 The Euler-Lagrange Equations

An even simpler method exists, we may re-write the generalized


equation of motion using the above relationship between generalized
force and gradient of the potential energy as
U d  T  T
    .
qk dt  qk  qk
(1.33)
Define the Lagrangian as
L  T U
sincewe have assumed holonomic constraints, we have U qk  0 . This
d  T  d  L 
allows the replacement of   with   giving
dt  qk  dt  q k 
d  L  L
  0
dt  qk  qk
(1.34)
This is the Euler-Lagrange equation, there is one for each generalized
coordinate qk.

Self Assessment Exercise C


1. Given U ( z)  mgz where z  b sin calculate the generalized force.
2. From the generalized equation of motion prove the Euler-
Lagrangian equation.

4.0 Conclusion

At the end we have been able to use D’Alembert’s principle to obtain a


closed set of equations for only the generalized coordinates; they are
known as the Euler-Lagrange equations. Writing these equations
explicitly only requires the knowledge of a scalar valued function: the
Lagrangian in terms of the generalized coordinates and the generalized
velocities.

30
PHY 301 CLASSICAL MECHANICS II

5.0 Summary

 D’Alembert’s Principle
D’Alembert’s Principle is
F 
 ( nc )  
i  p i  ri  0

where ri is a virtual displacement that is differential and satisfies
the constraints. D’Alembert’s principle may be re-written in
terms of generalized coordinates and forces

 ri
fk   p 
qk
 Generalized equation of motion
D’Alembert.s principle can be used to prove the generalized
equation of motion
d  T  T
fk     ,
dt  qk  qk
where T  T qk  , qk , t  is the kinetic energy written as a function
of the generalized coordinates.
 Euler-Lagrange’s Equation
When the non-constraint forces are conservative, they can be
written as gradients of a time independent potential,
 
F  iU rj  . From this, we can prove that the generalized

forces can also be written as gradients



fk   U (ql ).t  .
qk
If we then define the Lagrangian as L = T – U,
then, we can prove the Euler-Lagrangian equation
d  L  L
    0.
dt  qk  qk
In some cases, it is possible to write non conservative forces
using a potential function. If the non conservative force can be
written in terms of function U qk 
, qk  in the following manner;
U d  U 
fj   
q j dt  q j 
then one may include this function U as a potential energy in the
lagrangian and apply the Euler-Lagrangian equation.
If one has non-conservative forces that cannot be written in the
above form, one can still write down a generalized Euler-
Lagrangian equation
d  L  L
  f kno  L ,
 
dt  q j  q j

31
PHY 301 CLASSICAL MECHANICS II

where f kno L encompasses all forces that cannot be included in the


Lagrangian.

6.0 Tutor Marked Assignment (TMAs)

1. A pendulum of mass m hung on a rigid rod of length l whose (upper)


end is fixed. The mass is free to move on a circle of radius l on the
xz-plane. The angle that the rod makes with the vertical axis is ϴ
and the relation between r and ϴ is r  f ( )  l sin  ,0, l cos 
hence calculate (i) the Lagrangian and, (ii) the equation of motion.
2. Suppose now that this pendulum is free to move on the sphere
defined by the distance l from the anchor point (i.e., a two-
dimensional manifold) in which r is function of ϴ and ϕ, and are
related by r  f ( , )  l (sin cos , sin sin  , cos ) . Calculate the
Lagrangian and the corresponding equation of motion.

7.0 Further Reading and Other Resources

Hand, L. N. and Finch, J. D. (1998). Analytical Mechanics. Cambridge


University Press.
Thornton, S.T. and Marion, J. B. (2014.). Classical Dynamics of
Particles and Systems, Fifth Edition. Thomson Books/Cole
Goldstein, H., Poole, C., and Safko, J. (2000). Classical Mechanics,
Third Edition. Horoloma Book Jovanovich College Publisher.

Video Link:
https://www.youtube.com/playlist?list=PLtUBquogGkRukyAAWOmq2
LmQx8Yf9H_3E

32
PHY 301 CLASSICAL MECHANICS II

UNIT 2 LANGRANGE’S MECHANICS

CONTENTS

1.0 Introduction
2.0 Objectives
3.0 Main contents
3.1 Newton’s Second Law
3.1.1 Conservative Systems
3.1.2 Notation
3.2 Lagrange’s Equations
3.2.1 Generalized Coordinates
3.2.1 Lagrange’s Equations in Generalized Coordinates
3.3 Generalized Momenta
3.4 Lagrangian For Some Physical Systems
3.4.1 Example 1: 1-D motion—the pendulum
3.4.2 Example 2: 2-D motion in a central potential
3.4.3 Example 3: 2-D motion with time-varying
constraint
3.4.4 Example 4: Atwood Machine
3.4.5 Example 5: Elliptical wire
3.4.6 Example 6: Particle in e.m. field
3.5 Transformations of the Lagrangian
3.5.1 Point transformations
3.4.7 Gauge transformations
4.0 Conclusion
5.0 Summary
6.0 Tutor Marked Assignment (TMAs)
7.0 Further Reading and Other Resources

1.0 Introduction

It is not possible to develop the classical mechanical approach to


statistical mechanics without having some understanding of the
principles of classical mechanics. We will review the basic principles of
Newton’s Laws of Motion. We will also recast Newton’s second law
into the forms developed by Lagrange and Hamilton. These notes
provide some of the details about the Lagrangian and Hamiltonian
formulations of classical mechanics.

2.0 Objectives

By the end of this unit, you will be able to:


 express the Lagrangian L in Cartesian coordinates;
 transform L to generalized coordinates;
 give Lagrange’s equations in generalized coordinates.

33
PHY 301 CLASSICAL MECHANICS II

 Illustrate how to use Lagrange’s equations in generalized coordinates.


 Apply the approach to the free motion of a particle confined to
move on the perimeter of a ring.
 Solve related Problems.

3.0 Main Contents

3.1 Newton’s Second Law

We consider N particles moving in three-dimensional space, and we


describe the location of each particle using Cartesian coordinates. We let
mi be the mass of particle i, and we let xi, yi and zi be respectively the x,
y and z-coordinates of particle i. For time derivatives of the coordinates
(and all other physical observables), we use the ―dot‖ notation first
introduced by Isaac Newton
dxi d 2 xi
xi  and xi  .
dt dt 2
(2.1)
We let Fx be the x-component of the force on particle i. Then,
i

Newton’s Second Law takes the form Fx  mxi . i

For example, if we study the motion of a single particle of mass m


moving in one dimension in a harmonic potential with associated force
Fx=−kx, Newton’s second law takes the form  kx  mx . Solution of this
differential equation for the coordinate x as a function of the time t gives
a complete description of the motion of the particle; i.e. at any time t one
knows the location and velocity of the particle.

3.1.1 Conservative Systems

In quantum mechanics, we often restrict our attention to a class of


physical systems that are called conservative. In a qualitative sense
conservative systems are those for which the total energy E is the sum of
the kinetic and potential energies. For any isolated system E is
conserved (i.e. dE/dt = 0), and for conservative systems, the sum of the
potential energy and kinetic energy is conserved. Explicitly, we have the
following definition:
Definition: A classical mechanical system is conservative if there exists
a function U ( x1 , y1 , z1 , x2 ,... z N ) called the potential energy such that for
any coordinate xi (or yi or zi ) , we can write
U  U U 
Fxi   or Fyi   or Fzi  
xi   yi zi 
(2.2)
where Fx (or Fy or Fz ) is the x (or y or z) component of the force on
i i i

particle i.
34
PHY 301 CLASSICAL MECHANICS II

As an example, we can consider the one dimensional particle moving in


the harmonic well with force F  kx . For such a system, a potential
energy exists and is given by U ( x)  1 2 kx 2 . By differentiating the
potential energy with respect to x, the force is obtained. We can then be
sure that for a harmonic oscillator the total energy is conserved
1 2 1 2
E mx  kx .
2 2
(2.3)
3.1.2 Notation

The notation used for derivatives, we need to distinguish the explicit and
implicit dependence of a function on a variable. As an example, the
expression for the kinetic energy T of a system of N particles in
Cartesian coordinates is given by
T
1 N

2 i 1

mi x i2  y i2  z i2 
(2.4)
Equation 2.4 above for the kinetic energy is an explicit function of the
velocities of each particle; i.e. the xi , yi , zi  , but no other variables. We
can write
T
 m j x j .
x j
(2.5)
However, in the expression for T, there is no explicit dependence on the
time t or the coordinates xi , yi , zi , consequently, we write
T T
0 and 0
t z j
(2.6)

Equations 2.6 do not imply that the kinetic energy is independent of the
time or the z-component of the coordinate for particle j. Equation 2.6
only state that in the expression for T as written, there is no explicit
dependence of T on t or zj. If we want the actual (i.e. implicit)
dependence of the kinetic energy on time, we write
dT
which is not zero in the general case ( dT dt does equal zero for a free
dt
particle). We understand the notation  to represent a derivative of the
explicit dependence of a function as written on a variable and the
notation d to represent a derivative for the actual (i.e. implicit)
dependence of a function on a variable. It is important that the
differences between  and d are clear for the developments that follow.

35
PHY 301 CLASSICAL MECHANICS II

3.3 Lagrange’s Equations

We now derive Lagrange’s equations for the special case of a


conservative system in Cartesian coordinates. However, using
Lagrange’s equations in a more general coordinate systems will be
discussed after the derivation in Cartesian coordinates.
We begin with Equation 1.4 for the kinetic energy. We first differentiate
the expression with respect to one of the velocities.
T
 m j x j .
x j
We next take the implicit derivative of eq.1.4 above with respect to time
d T
 m j xj .
dt x j
(2.7)
The right hand side of equation 2.7 is the mass of the particle multiplied
by the acceleration of the particle, which by Newton’s law must be the
force on the particle. For a conservative system, We can express this
force by  U x j , so that
d T U
 .
dt x j x j
(2.8)
We now give a defining relation for the classical lagrangian:
Definition: The Classical Lagrangian is given by
L = T – U.
(2.9)
The classical Lagrangian is the difference between the kinetic and
potential energies of the system. Using this definition in eq. 1.7 we
obtain
d L L
  0.
dt x j x j
(2.10)

Equations 2.10 above are Lagrange’s equations in Cartesian coordinates.


We use the plural (equations), because Lagrange’s equations are a set of
equations. We have a separate equation for each coordinate xj. A
completely analogous set of equations is obtained for the other Cartesian
directions y and z.
We emphasize that Lagrange’s equations are just a new notation for
Newton’s second law. Example 2.1: Consider a one-dimensional
harmonic oscillator. The Lagrangian for the system is
1 2 1 2
L  T U  mx  kx
2 2
(2.11)

36
PHY 301 CLASSICAL MECHANICS II

d L L d
and Lagrange’s equation is   mx  kx  0 ,
dt x x dt
(2.12)
or mx  kx
which is just Newton’s Second law. To understand the utility of the new
notation, we need to introduce the notion of generalized coordinates.

Self Assessment Exercise A


Derive Lagrange’s equation of motion in Cartesian coordinates.

3.2.1 Generalized Coordinates


Much of Lagrange’s work was concerned with methods useful for
systems subject to external constraints.

figure 2.1
A simple example of the kind of problem that interested Lagrange is the
motion of a free particle of mass m confined to move on the perimeter of
a ring of radius R depicted in the figure above. Constraints on a
particle’s motion arise from some set of unspecified forces.

For the particle on a ring, it is possible to imagine some force of infinite


strength that limits the motion of the particle. The exact nature of the
force is not important to us. We only need to consider the confined
space.
For a particle on a ring, the Cartesian coordinates x and y are not the
most convenient to describe the motion of the particle. As a result of the
constraint, a single coordinate ϕ is sufficient to locate the particle. The
coordinate ϕ is defined to be the angle that a line connecting the current
location of the particle with the origin of coordinates makes with the x-
axis. The connections between ϕ and the Cartesian coordinates are given
by
x  R cos and y  R sin .
(2.13)

37
PHY 301 CLASSICAL MECHANICS II

The angle ϕ is an example of a generalized coordinate. Generalized


coordinates are any set of coordinates that are used to describe the
motion of a physical system. Cartesian coordinates and spherical polar
coordinates are other examples of generalized coordinates.
We may choose any convenient set of generalized coordinates for a
particular problem. For the particle in a ring example, the convenient
coordinate is ϕ. For systems with spherically symmetric potentials (the
motion of the earth about the sun, the hydrogen atom), we can choose
spherical polar coordinates. We label the i’th generalized coordinates
with the symbol qi, and we let qi represent the time derivative of qi.

Self Assessment Exercise B


1. Derive Lagrange’s equations of motions for a conservative system
in Cartesian coordinates.
2. Choose any convenient set of generalized coordinates for the
following systems: (i) free particle of mass m confined to
move on the perimeter of a ring of radius R depicted in the figure
2.1 above. (ii) the motion of a particle on a ring, (iii) for
systems with spherically symmetric potentials (the motion of the
earth about the sun, the hydrogen atom).

3.2.1 Lagrange’s Equations in Generalized Coordinates

Lagrange has shown that the form of Lagrange’s equations is invariant


to the particular set of
generalized coordinates chosen. For any set of generalized coordinates,
Lagrange’s equations
take the form
d L L
  0,
dt qi qi
(2.14)
exactly the same form that we derived in Cartesian coordinates. We
now illustrate how to use Lagrange’s equations in generalized
coordinates by applying the approach to the free motion of a particle
confined to move on the perimeter of a ring as discussed previously.
The meaning of the expression of ―free particle‖ is the absence of any
external forces. We can arbitrarily set the potential energy U to zero.
Then, in Cartesian coordinates, the Lagrangian for any free particle in
the xy-plane can be expressed as
1
L m( x 2  y 2 ) .
2
(2.15)
We now transform L to generalized coordinates using Eq.2.13. We need
the time derivatives of x and y expressed in terms of the generalized
coordinate system.

38
PHY 301 CLASSICAL MECHANICS II

x  R sin  R cos and y  R cos  R sin .


(2.16)
Owing to the constraint, R is a constant and R  0 , then
x  R sin and y  R cos
(2.17)
so that L becomes L
1
2
 
1
mR 22 cos2   sin 2   mR 22 .
2
(2.18)
Because of the constraint, the Lagrangian is a function of a single
coordinate ϕ. We finally give Lagrange’s equations
L d L L
 mR 2 , and  mR 2 and  0.
 dt  
(2.19)
So therefore,
d L d
 mR 2  mR 2  0

dt  dt
(2.20)
Equation 2.20 above implies the acceleration of the coordinates ϕ is zero
so that the particle moves with a constant generalized velocity 

Self Assessment Exercise C


1. Express Lagrange’s equation in generalized coordinates.
2. Derive the Lagrange’s equation for a free motion of a particle
confined to move on the perimeter of a ring in generalized
coordinates.

3.3 Generalized Momenta

Equation 2.20 can be interpreted to mean that the quantity


L   mR 2 is conserved. To fully explore the meaning of the
conservation of a quantity like L , consider the Lagrangian in
Cartesian coordinates for a particle of mass in one dimension
1 2
L mx  U ( x) .
2
(2.21)
By differentiating L with respect to the velocity x , we obtain the linear
momentum
L
 mx ,
x
(2.22)
which is conserved in the case of no external forces; i.e. the linear
momentum is conserved if U(x) is a constant. Using these simple
equation, we are lead to the following definition:

39
PHY 301 CLASSICAL MECHANICS II

Definition: The generalized momentum p i conjugate to the coordinate qi


is defined by
L
pi 
qi
(2.23)
for example in the particle in a ring, the generalized momentum
conjugate ϕ, L   p can be shown to be angular momentum of the
particle.

Self Assessment Exercise D


1. Define the generalized momentum in Cartesian and generalized
coordinates

3.4 Lagrangian for Some Physical Systems

3.4.1 Example 1: 1-D motion—the pendulum

One of the simplest nonlinear systems is the one-dimensional physical


pendulum (so called to distinguish it from the linearized harmonic
oscillator approximation). As depicted in Figure 2.2 below, the
pendulum consists of a light rigid rod of length l, making an angle ϴ
with the vertical, swinging from a fixed pivot at one end and with a bob
of mass m attached at the other end.

The constraint l = const and the assumption of plane motion reduces the
system to one degree of freedom, described by the generalized
coordinate ϴ. (This system is also called the simple pendulum to
distinguish it from the spherical pendulum and compound pendula,
which have more than one degree of freedom.)

40
PHY 301 CLASSICAL MECHANICS II

Figure 2.2 Physical Pendulums


The potential energy with respect to the equilibrium position ϴ = 0 is
U(ϴ) = mgl(1-cosϴ), where g is the acceleration due to gravity, and the
velocity of the bob is v  l , So that the kinetic energy
1 1
T mv  ml 2 2 .
2

2 2
The lagrangian, L= T – U, is thus
1
L( ,)  ml 2 2  mgl (1  cos ).
2
(2.24)
This is also essentially the Lagrangian for a particle moving in a
sinusoidal spatial potential, so the physical pendulum provides a
paradigm for problems such as the motion of an electron in a crystal
lattice or of an ion or electron in a plasma wave.
Lagrangian equation of motion is
d L L L L
 where  ml 2 and  mgl sin .
dt    
(2.25)
So therefore the Lagrangian equation of motion is
ml 2  mgl sin .
(2.26)

3.4.2 Example 2: 2-D motion in a central potential

Using plane polar coordinates, q  r ,  such that


x  r cos , y  r sin ,
(2.27)
so that
x  r cos  r sin y  r sin  r cos
(2.28)
1 1
hence the kinetic energy T  ( x 2  y 2 ) is found to be T  m(r 2  r 2 2 ) .
2 2
(2.29)
The lagrangian is
1
L m(r 2  r 2 2 )  U (r )
2
(2.30)
Since L is independent of ϴ then L   0 while L   mr 2
Therefore the ϴ-component of the Lagrangian equation of motion is
mr 2  0 while the r-component of the Lagrangian equation of motion
will be L r  mr while L r  mr2  U (r )
so the radial Lagrange’s equation of motion is
d
mr   mr2  U (r ) which gives
dt

41
PHY 301 CLASSICAL MECHANICS II

mr  mr2  U (r ) .
(2.31)

3.4.3 Example 3: 2-D motion with time-varying constraint

Consider a weight rotating about the origin on a frictionless horizontal


surface as depicted in the fig 2.3 below and constrained by a thread,
initially of length a, that is being pulled steadily downward at speed u
through a hole at the origin so that the radius r = a – ut.

figure 2.3
Then, on substituting for r in equation 2.29, the Lagrangian becomes,
L T 
1
2
 
m u 2  (a  ut ) 2 2 .
(2.32)
So we again have the conservation of angular momentum
m(a  ut)2  l  cons tan t
which can be integrated to give ϴ as a function of t,
  0  (l / mu)[1/(a  ut)  1/ a]  0  lt /[ma(a  ut)] .
(2.33)

3.4.4 Example 4: Atwood Machine

42
PHY 301 CLASSICAL MECHANICS II

Figure 2.4
Consider two weights of mass m1 and m2 suspended from a frictionless,
inertialess pulley of radius a by a rope of fixed length, as depicted in Fig.
2.4 above. The height of weight 1 is x with respect to the chosen origin
and the holonomic constraint provided by the rope allows us to express
the height of weight 2 as -x, so that there is only one degree of freedom
for this system.
The kinetic and potential energy are T
1
m1  m2 x 2 and
2
U  m1 gx  m2 gx . (2.34)
Thus, L  T  U
L
1
m1  m2 x 2  m1  m2 gx
2
(2.35)
and its derivatives are
L L
 m1  m2 g  m1  m2 x .
x x
(2.36)
The Lagrange’s equation of motion now becomes
m1  m2 x  m1  m2 g
m1  m2
 x  g.
m1  m2
(2.37)

43
PHY 301 CLASSICAL MECHANICS II

3.4.5 Example 5: Elliptical wire

For the elliptical wire of figure (1.1b), taking a and b to be constant, the
Lagragian is
L  T U 
2

m 2 2 2

a  sin   b 2 2 cos2   mgb sin  .
(2.38)
The Lagrange equations of motion is
d  L  d  T 
   
dt    dt   


d
dt
 
m a 2 sin 2   b 2 cos2  
   
 m a 2 sin 2   b 2 cos2   2m 2 a 2  b 2 sin  cos
(2.39)

L T U
 
  
 
 2m 2 a 2  b 2 sin  cos  mgb cos .
(2.40)
Thus
d  L  L
 
dt    
 
 0  m a 2 sin 2   b 2 cos2   mgb cos  0 .

(2.41)

3.4.6 Example 6: Particle in e.m. field

The fact that Lagrange’s equations are the Euler–Lagrange equations for
the extraordinarily simple and general Hamilton’s principle suggests that
Lagrange’s equations of motion may have a wider range of validity than
simply problems where the force is derivable from a scalar potential.
In particular, it is obviously of great physical importance to find a
Lagrangian for which Lagrange’s equations of motion equation 2.14
reproduce the equation of motion of a charged particle in an
electromagnetic field, under the influence of the Lorentz force,
mr  eE (r, t )  er  B (r, t ),
(2.42)
where e is the charge on the particle of mass m.
We assume the electric and magnetic fields E and B, respectively, to be
given in terms of the scalar potential Ф and vector potential A by the
standard relations
E     t A,
B    A.
(2.43)

44
PHY 301 CLASSICAL MECHANICS II

The electrostatic potential energy is eФ, so we expect part of the


1 2
Lagrangian to be mr  e , but how do we include the vector potential?
2
Clearly, we need to form a scalar since L is a scalar, so we need to dot A
with one of the naturally occurring vectors in the problem to create a
scalar.

The three vectors available are A itself, r and r . However we do not


wish to use A, since A·A in the Lagrangian would give an equation of
motion that is nonlinear in the electromagnetic field, contrary to
equation 2.42. Thus, we can only use r and r . Comparing equations
2.42 and 2.43 we see that r  A has the same dimensions as Ф, so let us
try adding that to form the total Lagrangian
1 2
L mx  e  er  A .
2
(2.44)
Taking q  q1, q2 , q3  x, y, z, we have
L  3 A j
 e   q j
qi qi j 1 qi
d L dA  A 3
A i 
and  mqi  e  mqi  e    q j .
dt q dt  t j 1 q j 
The lagrange’s equation of motion then becomes
  A i  e  A j A i 
mq  e     j
 e 
q  .
 qi t  j 1  qi q j 
(2.45)
This is simply equation 2.42 in Cartesian component form, so our
guessed Lagrangian is indeed correct.

Self Assessment Exercise E


1. Make sure you study and understand the above examples very well.

3.5 Transformations of the Lagrangian

3.5.1 Point transformations

Given an arbitrary Lagrangian L(q, q, t ) in one generalized coordinate


system, q  qi i  1, n(e.g. a Cartesian frame), we often want to know
the Lagrangian LQ, Q , t  in another coordinate system, Q  Qi i  1, n
(e.g. polar coordinates). Thus, suppose there exists a set g  gi i  1, n
of twice differentiable functions gi such that qi  gi Q,t , i = 1,…, n.
We require the inverse function of g also to be twice differentiable, in
which case g : Q  q is said to be C2 diffeomorphism. Note that we
have allowed the transformation to be time dependent, so
45
PHY 301 CLASSICAL MECHANICS II

transformations to a moving frame are allowed. The transformation g


maps a path in Q-space to a path in q-space. However, it is physically
the same path; all we have changed is its representation. What we need
in order to discuss how the Lagrangian transforms is a coordinate-free
formulation of Lagrangian dynamics. This is another virtue of
Hamilton’s principle i.e. S  0 on a physical path for all variations with
fixed endpoints and also since the action integral S given by
t2

S   dt Lq , q, t  is an integral over time only. Thus, if we can define L


t1

so that
 
L Q(t ), Q (t ), t  L(q(t ), q (t ), t )
(2.46)
for any path, then S is automatically invariant under the coordinate
change and will be stationary for the same physical paths, irrespective of
what coordinates they are represented in.
We can guarantee this trivially, simply by choosing the new Lagrangian
to be the old one in the new coordinates:
LQ, Q , t   Lg (Q, t ), g Q, Q , t , t  ,
(2.47)
gi n
g
where g i    Q j i , i = 1, …., n.
t i 1 Q j
One can prove that eq. (2.47) gives the correct dynamics by calculating
the transformation of the Euler–Lagrange equations explicitly and
showing that eq. 2.14 implies
d  L  L
   0,
dt  Q i  Qi
(2.48)
but clearly Hamilton’s principle provides a much simpler and more
elegant way of arriving at the same result, since equation 2.48 are
simply the Euler–Lagrange equations for S to be stationary in the new
variables.

3.4.7 Gauge transformations

It must be noted that we cannot always expect that the Lagrangian is of


the form T − V, but nevertheless the lagrangian gives a function for
which Lagrange’s equations represent the correct dynamical equations
of motion. Thus it is the requirement that the equations of motion be in
the form of Lagrange’s equations (or, equivalently, that Hamilton’s
principle apply) that is fundamental, rather than the specific form of L.
This naturally raises the question: for a given system, is there only one
Lagrangian giving the correct equations of motion, or are there many?
Clearly there is a trivial way to generate multiple physically equivalent
Lagrangians, and that is to multiply L by a constant factor. However, we

46
PHY 301 CLASSICAL MECHANICS II

usually normalize L in a natural way, e.g. by requiring that the part


linear in the mass be equal to the kinetic energy, so this freedom is not
encountered much in practice.
However, there is a more important source of non-uniqueness, known as
a gauge transformation of the Lagrangian in which L is replaced by L  ,
defined by
M M
L(q, q , t )  L(q, q , t )   q  .
t q
(2.49)

Example 2.2: Harmonic oscillator


Consider the harmonic oscillator Lagrangian, for a particle of unit mass
L 
1 2
2
x  02 x 2 
(2.50)
L L
where we take 0 to be constant. Since  x and   02 x, we
x x
immediately verify that Lagrange’s equation, equation 2.14 gives the
harmonic oscillator equation
x  02 x
(2.51)
1
Now add a gauge term, taking M  02 x . Then, from equation 2.49, the
2
new Lagrangian is
L 
2
x  2xx  02 x 2  .
1 2

(2.52)
L L 
Calculating  x  0 x, and   02 x   0 x, and substituting into
x x
the Lagragian equation of motion, we do indeed recover the harmonic
oscillator, equation, equation 2.51 and the new Lagrangian is a perfectly
valid one despite the fact that it is no longer in the natural form,T – U

Example 2.3: Electromagnetic gauge transformation


It is well known that the scalar and vector potentials in equation 2.43 are
not unique, since the electric and magnetic fields are left unchanged by
the gauge transformation
A (r , t )  (r , t )  A (r , t )   (r , t )
 (r , t )   (r , t )   (r , t )   t  (r , t ),
(2.53)
where  is an arbitrary scalar function (a gauge potential).
Using the gauge-transformed potentials   and A  to define new
Lagrangian L in the same way as L was defined by equation 2.44 we
have

47
PHY 301 CLASSICAL MECHANICS II

1 2
L  mx  e  er  A .
2
(2.54)
Substituting equation 2.53 in equation 2.54 we find
(e )
L  L   r  (e )
t
(2.55)
which is exactly of the form equation 2.49 with M  e . Thus
electromagnetic and Lagrangian gauge transformations are closely
related.

Self Assessment Exercise F


1. State what is meant by point and gauge transformation.
2. Find the gauge transformation of Lagrangian of harmonic oscillator
and electromagnetic fields.

4.0 Conclusion

The Classical Lagrangian is given by


L = T – U.
(2.9)
The classical Lagrangian is the difference between the kinetic and
potential energies of the system. Lagrange has shown that the form of
Lagrange’s equations is invariant to the particular set of generalized
coordinates chosen. For any set of generalized coordinates, Lagrange’s
equations take the form
d L L
  0.
dt q i qi

5.0 Summary

 Using the ―dot‖ notation introduced by Isaac Newton


dx d 2 xi
xi  i and xi  2
dt dt
 For motion of a single particle of mass m moving in one
dimension in a harmonic potential with associated force Fx=−kx,
Newton’s second law takes the form  kx  mx
 A classical mechanical system is conservative if there exists a
function U ( x1 , y1 , z1 , x2 ,... z N ) called the potential energy such that
for any coordinate xi (or yi or zi) we can write
U  U U 
Fxi   or Fyi   or Fzi   where
xi   yi zi 
Fxi (or Fyi or Fzi ) is the x ( or y or z) component of the
force on particle i.

48
PHY 301 CLASSICAL MECHANICS II

 The Classical Lagrangian is given by


L=T–U
 The classical Lagrangian is the difference between the kinetic and
potential energies of the system. Using eq. 1.7 we obtain
d L L
  0.
dt x j x j
 For any set of generalized coordinates, Lagrange’s equations take
the form
d L L
  0, .
dt qi qi
 The generalized momentum pi conjugate to the coordinate qi is
defined by
L
pi  .
qi
 Given an arbitrary Lagrangian L(q, q, t ) in one generalized
coordinate system, q  qi i  1, n (e.g. a Cartesian frame), we
often want to know the Lagrangian LQ, Q , t  in another
coordinate system, Q  Qi i  1, n (e.g. polar coordinates).
However, there is a more important source of non-uniqueness,
known as a gauge transformation of the Lagrangian in which L is
replaced by L  , defined by
M M
L(q, q , t )  L(q, q , t )   q  .
t q

6.0 Tutor Marked Assignments (TMAs)

1. As a model of the motion of a fluid element or dust particle in a


planetary (e.g. Earth’s) atmosphere, consider the motion of particle
of unit mass constrained to move on the surface of a perfectly
smooth sphere of radius R rotating with angular velocity ω about
the z-axis. Suppose the force on the particle is given by an effective
potential U(ϴ,ϕ), where ϴ and ϕ are the latitude and longitude
respectively.
(a) Write down the Lagrangian in a frame rotating with the planet,
taking the generalized coordinates to be the latitude and longitude so
that z  R sin  , x  R cos cos(  t ) , y  R cos sin(  t ) , where
x, y, z is a non-rotating Cartesian frame. (b) Write down the
equations of motion, and find a first integral (i.e. constant of the
motion) in the case where U is independent of longitude. Among
this class of potentials find the special case U  U ( ) required to
make equilibrium possible (i.e. so that the equations of motion
admit the solution     0 at each latitude).
Answers

49
PHY 301 CLASSICAL MECHANICS II


x   R sin  cos   t   R    cos sin   t  

y   R sin  sin   t   R    cos cos   t  
z  R cos
Thus the kinetic energy is given by
T
2

1 2 2 2
x  y  z 
1
2
  2
    
 R 2  2 sin 2  cos2   t   sin 2   t      cos2  sin 2   t   cos2   t    2 cos2  
1
  
 R 2  2     cos2 
2
2

and the Lagrangian, L = T – U, by
L
1 2 2
2
  
R      cos2   U ( , ) .
2

L U
L


  R 2    sin  cos 
2

U

, and



L
L
Thus   R 2 and
 
 R 2    cos2   
d L L
and thus the lagrange equations of motion  and
dt  
d L L
 become
dt  

R 2   R 2    sin  cos 
2
 U

and


R 2cos2   2 R 2    sin cos   
U

.
In the case where U = U(ϴ), ϕ is an ignorable coordinate and thus
L L

given by   R 2 (  ) cos2  is an integral of the motion.
 
If U0 be such that     0 for all ϴ and ϕ, and use the identity
U 0 1 2 2
sin 2  2 sin cos , then,  R  sin 2
 2
Or integrating, we have
1 2 2 1
U 0 ( )  R  cos 2  cos 2  const  R 2 2 cos2 
4 2
where the second equality follows fom the identity
1 2
cos 2  2 cos2   1 and the choice const  R .
4

2. Consider a charged particle constrained to move on a non-rotating


smooth insulating sphere, immersed in a uniform magnetic field B =
Bez, on which the electrostatic potential is a function of latitude and
longitude. Write down the Lagrangian in the same generalized

50
PHY 301 CLASSICAL MECHANICS II

coordinates as above and show that it is the same as for the particle
on the rotating planet with appropriate identifications of ω and U.

The Lagrangian is L  T  er  A  e, with ω set o zero. We have now


need a vector potential A such that B z   x A y   y A x  B . A suitable

eB xy  yx 
1 1 1
choice is A x   B y, A y   B x, A z  0 . Thus, er  A 
2 2 2
1
er  A 
2
   
eB R 2 cos cos   sin  sin    cos cos  sin   sin  cos   cos sin  
1
 eB R 2 cos2 
2

Thus, the Lagrangian is


L
1
2
 1
2

mR 2  2   2 cos2   eB R 2 cos2   e ,   .

Completing the square we can write the Lagrangian in the form


   eB 
2
 e2B 2 R 2
cos2   e ,   .
1 2 2
L  mR      cos   
2

2   2m   8m
This is the same as the Lagrangian for particle on the rotating planet
where
eB e2B 2 R 2
 and U ( ,  )  cos2   e  ,   Note that the equation
2m 8m
1 eB
above can also be written as   0 , where  0  is the
2 m
cyclotron frequency.
3. Write down a Lagrangian for the problem of two particles of mass
m1 and m2 connected by a light rigid rod of length l in a
gravitational field g. Take the generalized coordintes of the system
to be q  x, y, z, , , with the coordinates of two particles being
given by
x1  x   1l sin  cos ,
y1  y   1l sin  sin  ,
z1  z   1l cos ,
x 2  x   2 l sin  cos ,
y 2  y   2 l sin  sin  ,
z 2  z   2 l cos ,
m2 m1
where 1  and  2  (so that (x, y, z) is the
m1  m2 m1  m2
centre of mass).

51
PHY 301 CLASSICAL MECHANICS II

Answer
x1  x   1l sin  sin    1l cos cos
x  x   l sin  sin    l cos cos .
2 2 2

Because m11  m2 2 the cross terms cancel when we expand the
x-contribution to the kinetic energy.
m  m2 2
1
2
1
m1 x12  m2 x22  1
2 2
x 
m1m2l 2
2(m1  m2 )

 sin sin    cos cos 
2

Similarly
m  m2 2
1
2
1
m1 y12  m2 y 22  1
2 2
y 
m1m2l 2
2(m1  m2 )

 sin cos   cos sin   2

1 1 m  m2 2 m1m2l 2  2 2
m1 z12  m2 z22  1 z   sin  .
2 2 2 2(m1  m2 )
Thus, adding the kinetic energy, we have
m1  m2 2
T
2
 
x  y 2  z 2 
m1m2
2(m1  m2 )

 2  sin 2  2 
The potential energy is given by
U  m1 z1  m2 z2 g  m1  m2 zg
and the Lagrangian is L = T – U
m1  m2 2
L
2
( x  y 2  z 2 ) 
m1m2
2(m1  m2 )
 
 2  sin 2 2  m1  m2 zg

7.0 Further Reading and Other Resources

Hand, L. N. and Finch, J. D. (1998). Analytical Mechanics. Cambridge


University Press.

Thornton, S.T. and Marion, J. B. (2014.). Classical Dynamics of


Particles and Systems, Fifth Edition. Thomson Books/Cole

Goldstein, H., Poole, C., and Safko, J. (2000). Classical Mechanics,


Third Edition. Horoloma Book Jovanovich College Publisher.

Video Link:
https://www.youtube.com/playlist?list=PLtUBquogGkRukyAAWOmq2
LmQx8Yf9H_3E

52
PHY 301 CLASSICAL MECHANICS II

UNIT 3 HAMILTONIAN MECHANICS

CONTENTS

1.0 Introduction
2.0 Objectives
3.0 Main Contents
3.1 Legendre Transform
3.2 The Classical Hamiltonian and Hamilton’s Equations
3.2.1 Derivation of Hamiltonian and Hamilton’s equation
of motion if the Lagrangian is a function of two
variables q and z (and possibly time).
3.2.2 Example 1: Scalar potential
3.2.3 Example 2: Physical pendulum
3.3 Construction of the Hamiltonian in Spherical Polar
Coordinates - Central Force Motion
4.0 Conclusion
5.0 Summary
6.0 Tutor Marked Assignments
7.0 Further Reading and Other Resources

1.0 Introduction

Hamiltonian mechanics is a reformulation of classical mechanics that


was introduced in 1833 by Irish mathematician William Rowan
Hamilton. It arose from Lagrangian Mehanics, a previous reformulation
of classical mechanics introduced by Joseph Louis Lagrange in 1788,
but can be formulated without recourse to Lagrangian mechanics using
symplectic spaces. [The Hamiltonian method differs from the
Lagrangian method, in that, instead of expressing second-order
differential constraints on an n-dimensional coordinate space (where n is
the number of degrees of freedom of the system)], it expresses first-
order constraints on a 2n-dimensional phase space.

As with Lagrangian mechanics, Hamilton's equations provide a new and


equivalent way of looking at classical mechanics. Generally, these
equations do not provide a more convenient way of solving a particular
problem. Rather, they provide deeper insights into both the general
structure of classical mechanics and its connection to quantum
mechanics as understood through Hamiltonian mechanics, as well as its
connection to other areas of science.

53
PHY 301 CLASSICAL MECHANICS II

2.0 Objectives

By the end of this unit, you will be able to:


 Explain Legendre transform.
 Understand the application of legendre transform in
thermodynamics.
 Find the Legendre transform of any function.
 Derive the Hamiltonian from Legendre transform and the
corresponding Hamilton’s equation of motion.
 Derive the Hamiltonian and Hamilton’s equation of motion in
spherical polar coordinates.
 Solve related Problems.

3.0 Main Contents

3.1 Legendre Transform

The Legendre transform is a method of changing the dependence of a


function of one set of variables to another set of variables.
In mathematics, it is often desirable to express a functional relationship
f(x) as a different function, whose argument is the derivative of f , rather
df
than x . If we let p  be the argument of this new function, then this
dx
new function is written f * ( p) and is called the Legendre transform of
the original function, named after Adrien-Marie Legendre. The
Legendre transform f * of a function f is defined as follows:
f * ( p)  max x ( px  f ( x))
The notation maxx indicates the maximization of the expression with
respect to the variable x while p is held constant. The Legendre
transform is its own inverse. Like the familiar Fourier transform, the
Legendre transform takes a function f(x) and produces a function of a
different variable p. However, while the Fourier transform consists of an
integration with a kernel, the Legendre transform uses maximization as
the transformation procedure. The transform is well behaved only if is a
convex function.
The Legendre transformation is an application of the duality relationship
between points and lines. The functional relationship specified by f(x)
can be represented equally well as a set of (x, y) points, or as a set of
tangent lines specified by their slope and intercept values.
The Legendre transformation can be generalized to the Legendre-
Fenchel transformation. It is commonly used in thermodynamics and in
the Hamiltonian formulation of classical mechanics.

54
PHY 301 CLASSICAL MECHANICS II

figure 1.1 The Legendre transformation: the cuve graph is f(x).


Its tangent with slope P intersects the y-axis at the point –f*(p).

The Legendre transform is most often used in the study of


thermodynamics. Recall from thermodynamics, there are two free
energy functions called the Gibbs free energy and the Helmholtz free
energy. The total differential of the Helmholtz free energy is given by
dA  - SdT - PdV ,
(3.1)
where S is the entropy, T is the temperature, V is the volume and p is the
pressure. From the total differential, it is evident that the Helmholtz free
energy is expressed as a function of the temperature and the volume; i.e.
A = A(T, V ). Now suppose we prefer to express the state of our system
in terms of temperature and pressure rather than temperature and volume.
We can define a new function G by
G  A  PV
(3.2)
where we have added to A the product of the variable we want (p) and
the variable we want to eliminate (V ). The algebraic sign of the
included PV product is chosen to be the opposite of the algebraic sign of
the PdV term in Equation 3.1. Taking the differential of the expression
for G we obtain
dG  dA  PdV  V dP  - SdT - PdV  PdV  V dP  - SdT  V dP .
(3.3)
It is evident that G is a function of T and P as desired. Of course, G is
the Gibbs free energy, and G is said to be the Legendre transform of A.
As previously mentioned, in Equation 3.2 the product PV is included
with an algebraic sign opposite to the sign of PdV in Equation 3.1, so
that the cancellation of the two PdV terms is assured.

Self Assessment Exercise A


Find the Legendre transform of the function ex.

55
PHY 301 CLASSICAL MECHANICS II

Show that the Legendre transform satisfies the following algebraic


properties.
1. f ( x)  ag( x)  f * ( p)  ag * ( p / a)
2. f ( x)  g (ax)  f * ( p)  g * ( p / a)
3. f ( x)  g ( x  b)  f * ( p)  g * ( p)  b
4. f ( x)  g 1 ( x)  f * ( p)   pg * (1 / p) .
where f * ( p) is the Legendre transform

3.2 The Classical Hamiltonian and Hamilton’s Equations

We now apply the notion of the Legendre transform to the classical


Lagrangian. In our previous developments, we have taken L to be a
function of all the generalized coordinates and their respective time
derivatives; i.e. L  Lqi 
, qi , t  . For generality, we have also included
the possibility that the Lagrangian has explicit time dependence. Such
explicit time dependence can occur when the external forces acting on a
system are time-dependent.

The resulting time-dependent potentials can be important in systems, as


for example, the study of the interaction of radiation with matter. Light
is composed of electric and magnetic fields that oscillate in time, and
when light interacts with matter, the electrons are subjected to time-
dependent potentials. The quantum treatment of spectroscopy includes
time dependent potentials, and we generalize the Lagrangian to admit
such time dependences.

We now use the Legendre transform to define a new function where we


replace the velocity (the qi ) dependence by a dependence on the
generalized momenta. The transformed function is as follows:
Definition: For a system of particles each having masses mi described by
a set of generalized coordinates qi, the classical Hamiltonian is defined
by
H   pi qi  Lqi , qi , t  .
i
(3.4)
As we now show, the particular choice of the relative signs of the first
and second terms in Equation 3.4 above makes the classical Hamiltonian
a natural function of the generalized coordinates and momenta rather
than the generalized coordinates and the velocities. The reason that the
sum is included with a positive sign and the Lagrangian is included with
a negative sign (rather than the opposite) is made clear shortly when we
identify the meaning of the classical Hamiltonian.
We now take the total differential of equation 3.4

56
PHY 301 CLASSICAL MECHANICS II

 L L  L
dH    pi dqi  qi dpi  dqi  dqi   dt
i  qi qi  t
(3.5)
The derivative L qi is the definition of the generalized momenta pi.
From Lagrange’s equation, we can write
d L L
pi  0 or p i  .
dt qi qi
(3.6)
Then the total differential of the classical Hamiltonian becomes
L
dH   ( pi dqi  qi dpi  p i dqi  pi dq )  dt
i t
L
  qi dpi  p i dqi   dt.
t
(3.7)
The classical Hamiltonian is manifestly a function of the generalized
coordinates and momenta rather than the generalized coordinates and
velocities from equation 3.7, we have
H H H L
 qi ,   p i and  .
pi qi t t
(3.8)
Equations 3.8 are called Hamilton’s equation of motion.
Consider the motion of a particle of mass in one dimension with the
Lagrangian given in equation 2.21. The generalized momentum is given
as px  L x  mx, and from the definition of the Hamiltonian, we have
1  1
H  px x  L  mxx   mx 2  U ( x)   mx  U ( x) ,
2  2
(3.9)
the Hamiltonian is seen to be the sum of the kinetic and potential
energies of the system.
For conservative system, H is the total energy. But the equation above is
not acceptable because H is not written explicitly as a function of the
generalized momentum. We must substitute x  p x m then equation 3.9
above becomes
px2
H  U ( x) .
2m
(3.10)
The Hamiltonian is then seen to be an expression for the total energy of
a conservative system in terms of the generalized coordinates and
momenta. With the Hamiltonian expressed in terms of the proper
variables, we can give Hamilton’s equations of the system as
H p H U
 x  x and    p x .
px m x x
(3.11)

57
PHY 301 CLASSICAL MECHANICS II

3.2.1 Derivation of Hamiltonian and Hamilton’s equation of


motion if the Lagrangian is a function of two variables q and
z (and possibly time).

Consider the Legendre transformation of the Lagragian with respect to


the variables z.
L * (q, p, t )  pz(q, p, t )  L(q, z(q, p, t ), t ),
L
where z(q, p, t ) is the inverse relation to p  (q, z, t ), i.e. it is our
z
definition of the conjugate momentum. The Legendre transformation of
the Lagragian is called the Hamiltonian, and it is commonly denoted by
the letter H,
H (q, p, t )  pz(q, p, t )  L(q, z(q, p, t ), t ).
We now examine the derivatives of H with respect to its arguments.
Firstly
H z L L z
( q, p, t )  p ( q , p, t )  (q, z (q, p, t ), t )  (q, z (q, p, t ), t ) (q, p, t ).
q q q z q

L
However, the first and third term cancel, since p  (q, z, t ), so that
z
H L
(q, p, t )   (q, z (q, p, t ), t )
q q
Secondly
H L z
(q, p, t )  z (q, p, t )  p (q, z (q, p, t ), t ) (q, p, t ).
p p p
Also, the second and the third term cancel to become
H
( q , p, t )  z ( q , p , t )
p
The Hamilton’s equations are
dq H dp H
 ( q, p, t ) ,  ( q , p, t ) ,
dt p dt p
this procedure can be extended to vector-valued q. Given the
Lagrangian L(q, z, t ) , we define the Hamiltonian as
n
H (q, p, t )   pi z i (q, p, t )  L(q, z (q, p, t ), t ) .
i 1

Where the transformation between (q,z,t) and (q,p,t) is


L
pi  ( q, z , t ) .
z i
Hamilton’s equations are system of 2n first-order equations
dqi H dp H
 ( q, p, t ) ,  ( q, p, t )
dt pi dt qi

58
PHY 301 CLASSICAL MECHANICS II

3.2.2 Example 1: Scalar potential

Consider the Lagrangian for a particle in Cartesian coordinates, so q ={x,


y, z} may be replaced by r = xex +yey +zez. Also assume that it moves
under the influence of a scalar potential U (r, t) so that the natural form
of the Lagrangian is
1
L  T U  m r  U (r , t )
2

2
(3.12)
L
General momentum p   mr ,
r
(3.13)
so that in this case the canonical momentum is the same as the ordinary
kinematics’ momentum. Equation.3.8 is solved trivially to give q  u( p)
where u( p)  p m . Thus, from eq.3.9 we have
p  p2
2
 p2
H    U (r , t )    U (r , t ),
m  2m  2m
 
 T  U.
(3.14)
That is, the Hamiltonian is equal to the total energy of the system,
kinetic plus potential. The fact that the Hamiltonian is an important
physical quantity, whereas the physical meaning of the Lagrangian is
more obscure, is one of the appealing features of the Hamiltonian
approach. Both the Lagrangian and Hamiltonian have the dimensions of
energy, and both approaches can be called energy methods. They are
characterized by the use of scalar quantities rather than the vectors
encountered in the direct use of Newton’s second law.
An example is the harmonic oscillator, Hamiltonian corresponding to
the Lagrangian equation 3.9 is
p 2 m0 x 2
H  .
2m 2
(3.15)
From equation 3.9 the Hamiltonian equations of motion are
p
x  , p  m0 x .
m
(3.16)
Gauge-transformed Harmonic Oscillator.
Now consider the gauge-transformed harmonic oscillator Lagrangian
mx 2  20 xx  02 x 2 
1
L 
2
(3.17)
The canonical momentum is thus

59
PHY 301 CLASSICAL MECHANICS II

L
p  m( x  0 x)
x
(3.18)
and we see that the gauge transformation has affected a transformation
of the canonical momentum. Even though the generalized coordinate x
remains the same.
Solving equation 3.9 for x , we find u ( p )  ( p  m 0 x . Hence
m

H  p
 p  m0 x 
2
1
 m02 x 2  T  U
m 2
(3.19)
Thus, even though L was not of the natural form T – U in this case, the
Hamiltonian remains equal to the total energy, thus confirming that it is
a quantity with a more direct physical significance than the Lagrangian.
(The functional form of the Hamiltonian changes under the gauge
transformation because the meaning of p changes).

3.2.3 Example 2: Physical pendulum

A nonlinear one-dimensional case is provided by the physical pendulum


was introduced in module 2 The Hamiltonian is
H  p  L
(3.20)
becomes
p2
H ( , p )   mgl (1  cos )
2ml 2
(3.21)
which again is of the form T + U Thus the Hamiltons equations are
p
p  ml 2, and   2 .
ml
(3.22)

Self Assessment Exercise B


1. Give the expression for the Hamiltonian of a system in generalized
coordinates.
2. Write the Hamiltonian equation of motion.
3. Derive the Hamiltonian of Scalar potential and physical pendulum
and their corresponding Hamilton’s equation of motion.

60
PHY 301 CLASSICAL MECHANICS II

3.3 Construction of the Hamiltonian in Spherical Polar


Coordinates - Central Force Motion

figure 3.1: systems with spherical symmetry (spherical polar


coordinates)

For systems with spherical symmetry (e.g. the rigid rotator), spherical
polar coordinates are the most convenient set of generalized coordinates.
As depicted above, the spherical polar coordinates are r, ϴ and ϕ. The
coordinate r is the distance from the origin of coordinates to the particle,
ϴ is the angle the line connecting the origin of the coordinates to the
particle position makes with the z-axis and ϕ is the angle the projection
of the line defining ϴ onto the xy-plane makes with the x-axis. The
connections between Cartesian coordinates and spherical polar
coordinates can be derived readily using trigonometry. The result is
x  r sin cos y  r sin sin z  r cos .
(3.23)
Another important relation is the direct result of the Pythagoras theorem
r  x2  y2  z 2 .
(3.24)
Consider a particle of mass m moving in three-dimensional space to a
conservative central force with associated potential energy U(r). The
meaning of a central force is the potential energy is a function only of
the coordinate and independent of ϴ and ϕ. In Cartesian coordinates, the
Lagrangian for the system is
L
1
2
 
m x 2  y 2  z 2  U .
(3.25)

61
PHY 301 CLASSICAL MECHANICS II

To transform L from Cartesian to spherical polar coordinates, we need


expression for the time derivatives of each Cartesian coordinates in
terms of spherical polar coordinates.
Using equation 3.23 we have
x  r sin cos  r cos cos  r sin sin
(3.26)
y  r sin sin  r cos sin  r sin cos
(3.27)
and z  r cos  r sin
(3.28)
We then substitute equations 3.26-3.28 into eq.3.25. After some algebra,
the result is
L
1
2
 
m r 2  r 2 2  r 2 sin 2 2   U (r ) .
(3.29)
The lagrange equation is
d dU
mr  mr 2  mr sin 2 2   0,
dt dr
(3.30)
for the ϴ- coordinate, we have
d
dt
 
mr 2  mr 2 sin cos2  0
(3.31)
and for the ϕ-coordinate we have
d
dt
mr 2 sin 2    0 .

(3.32)
The equation for the ϕ-coordinate is an expression of the conservative of
the momentum conjugate to the coordinate ϕ (the angular momentum).
To construct the classical Hamiltonian, we need expressions for the
generalized momenta conjugate to each of the spherical polar
coordinates. These expressions have already been obtained when
constructing Lagrange’s equations. In particular,
L L L
pr   mr, p   mr 2, and p   mr 2 sin 2 .
r  
(3.33)
Then using the definition of the classical Hamiltonian
H  p r r  p   p   L


1
2
 
m r 2  r 2 2  r 2 sin 2  2  U (r ).
(3.34)
Finally equation 3.34 above must be transformed to replace the
velocities with generalized momenta. We finally obtain

62
PHY 301 CLASSICAL MECHANICS II

2
p2 p2 p
H  r  2  U (r ) .
2m 2mr 2mr sin 2 
2

(3.35)
If needed the equations of motion can be obtained by applying
Hamilton’s equations to the constructed Hamiltonian.

Self Assessment Exercise C


1. Derive the Hamiltonian in spherical polar coordinates.
2. Derive the corresponding equation of motion in spherical polar
coordinates.

4.0 Conclusion

The Legendre transform is a method of changing the dependence of a


function of one set of variables to another set of variables. It is
commonly used in thermodynamics and in the Hamiltonian formulation
of classical mechanics.
Using Legendre transform to classical mechanics then, for a system of
particles each having masses mi described by a set of generalized
coordinate’s qi, the classical Hamiltonian is defined by
H   pi qi  Lqi , qi , t  .
i

The classical Hamiltonian is manifestly a function of the generalized


coordinates and momenta rather than the generalized coordinates and
velocities.
Hamilton’s equation of motion are.
H H H L
 qi ,   p i and  .
pi qi t t

The Hamiltonian is then seen to be an expression for the total energy of


a conservative system in terms of the generalized coordinates and
momenta. With the Hamiltonian expressed in terms of the proper
variables, we can give Hamilton’s equations of the system
H p H U
 x  x and    p x
px m x x

5.0 Summary

 The Hamiltonian and Hamilton’s equations, the Hamiltonian


function is derived from the Lagrangian function via the
lengendre transformation
H   pk qk  L ,
k

H  H qk 
, pk , t  .

63
PHY 301 CLASSICAL MECHANICS II

 The Hamiltons equation of motion are


H H dH H L
q k  ,  p k  ,   .
pk q k dt t t

6.0 Tutor Marked Assignments (TMAs)

1. Use the Lagrangian to construct the Hamiltonian for the system.


Answer
m1m2
Let M  m1  m2 be the total mass and   that is, the
m1  m2
reduced mass. Then,
px2 p2 p2 p2 p2
H   Y  z  2   Mgz .
2M 2M 2M 2l 2l 2 sin 2 

2. Find the Hamiltonian corresponding to the Coriolis


3. Lagrangian
L
1 2 
2
  
R   cos2      U ( , ) .
2

Answer
From the above,
L p
 R 2 , so that   2 .
p 
  R
L
Also p    R 2 cos2    

  p
so   2  2  
R cos 
H  p   p   L
 

p 2
p2 p2 p2
 2  2   p    U ( ,  )
R R cos2  2 R 2 2 R 2 cos2 
p2 p2
    p  U ( ,  )
2 R 2 2 R 2 cos2 

4. Find the Hamiltonian corresponding to the Lagrangian of an


harmonic oscillator of problem
Answer
1
The Lagragian L  mx 2  U ( x), which is the standard form treated.
2
p
Thus, x  and the Hamiltonian is
m
p 2 m02  2 x 4 
H   x  2  .
2m 2  l0 

64
PHY 301 CLASSICAL MECHANICS II

5. Consider the motion of a particle of charge e and mass m in a


straight infinitely long magnetic confinement system with vector
potential A   ( x, y )e z .

If the Hamiltonian is
px2 p y2  px  e x, y 2
H    ,
2m 2m 2m
write down the Hamiltonian equation of motion

Answer
The Hamiltonian equation of motion are
H p
x   z,
p x m
H py
y   ,
p y m
H  p z  e 
z   ,
p z m
and
H e p z  e  
p x    .
x m x
H e p z  e  
p y    .
y m y
H
p z    0.
z
6. Suppose a system has a Lagrangian
q 22
Lq1 , q 2 , q1 , q 2   q12   k1 q12  k 2 q1 q 2 .
a  bq12
Find the equations of motion using the Hamiltonian formulation.

7.0 Further Reading and Other Resources

Hand, L. N. and Finch, J. D. (1998). Analytical Mechanics. Cambridge


University Press.

Thornton, S.T. and Marion, J. B. (2014.). Classical Dynamics of


Particles and Systems, Fifth Edition. Thomson Books/Cole

Goldstein, H., Poole, C., and Safko, J. (2000). Classical Mechanics,


Third Edition. Horoloma Book Jovanovich College Publisher.

Video Link:
https://www.youtube.com/playlist?list=PLtUBquogGkRukyAAWOmq2
LmQx8Yf9H_3E

65
PHY 301 CLASSICAL MECHANICS II

MODULE 3 CENTRAL FORCE AND SCATTERING

Unit 1 The Generic Central Force Problem


Unit 2 Kepler’s Problem
Unit 3 Scattering Cross Section

UNIT 1 THE GENERIC CENTRAL FORCE PROBLEM

CONTENTS

1.0 Introduction
2.0 Objectives
3.0 Main Contents
3.1 The Generic Central Force Problem
3.1.1 The Equation of Motion
3.1.2 Reduction to a one body problem
3.2 Dynamics of an Isolated Two-Body Central-Force System
3.2.1 Reduction to one dimension
3.3 The formal Solution To The Equation Of Motion
3.3.1 Integration of Equation of motion
3.3.2 A Differential Relation between r and ϴ
3.3.3 Qualitative Dynamics
4.0 Conclusion
5.0 Summary
6.0 Tutor Marked Assignment
7.0 Further Reading and Other Resources

1.0 Introduction

The problem of the motion of two bodies interacting via a central force
is an important application of Lagrangian dynamics, and the
conservation theorems we have learned about. Central forces describe a
large variety of classical systems, ranging from gravitationally
interacting celestial bodies to electrostatic and nuclear interactions of
fundamental particles. The central force problem provides one of the
few exactly solvable problems in mechanics. And central forces
underlying most scattering phenomena, again ranging from gravitational
to electrostatics to nuclear.

66
PHY 301 CLASSICAL MECHANICS II

2.0 Objectives

By the end of this unit, you will be able to:


 Define central force and know the properties of an isolated two
body central force system.
 Discuss the reduction of the two body problem to a
mathematically equivalent problem of a single particle moving in
one direction.
 Discuss the reduction of a two-body three-dimensional problem
to one with single degree of freedom.
 Explain different shape of the effective potential energy function
and its implications for the motion of the system.
 Give the differential relation between r and ϴ.
 Solve related Problems.

3.0 Main Content

3.1 The Generic Central Force Problem

We first discuss the central force problem in general terms, considering


arbitrary radially dependent potential energy functions.

3.1.1 The Equation of Motion

Review of Central Forces.

Central force is defined as one that satisfies the strong form of Newton’s
third law. That is, given two particles a and b, the force exerted by
particle a on b is equal and opposite to that exerted by particle b on
particle a, and, moreover, the force depends only on the separation of the
two particles and points along the vector between the two particles.
Mathematically, this means,
  
f ab   f ba , f ab  f ab (rab )rˆab

    rab
where rab  ra  rb , rab  rab , rˆab  .
rab
(1.1)

Properties of an Isolated Two-Body Central-Force System


Let us now consider an isolated two-body system interacting via
a conservative central force. There are no other forces acting on the
bodies.
From the concepts of force, momentum, and energy for systems of
particles. One can abstract the following facts about isolated two-body
system:

67
PHY 301 CLASSICAL MECHANICS II

• Since no external forces act on the system, Newton’s second law for
systems of particles tells us that the total linear momentum is conserved.
d  d  
0 p  (ma ra  mb rb ) . (1.2)
dt dt
  
Since P is constant, the velocity of the center-of-mass R  P / M is
fixed and thus the center-of-mass system is inertial. We may therefore
 
assume, without loss of generality, that ra and rb are coordinates in the

center of mass system, where P vanishes and the center of mass is at the
origin:
  
0  R  ma ra  mb rb .
  
0  P  ma ra  mb rb .
(1.3)
This eliminates three of the six degrees of freedom in the problem. The

difference coordinate rab is now
    m   m 
rab  ra  rb  ra 1  a   rb 1  b  .
 mb   ma 
(1.4)

Defining the reduced mass


ma mb

ma  mb
(1.5)
gives us
     
ra  rab rb  rab .
ma mb
(1.6)
We shall see that the dynamics of the two-particle system will be
reduced to that of a single particle with mass μ moving in the potential
U(rab). We may consider two simple limits immediately:
 In the limit when mb is far greater than ma we have   ma ,
  
rb  o, and rab  ra . That is, the center of mass is fixed on the
heavier mass and the motion is entirely of the smaller mass.
 
m  rab  rab
 In the limit ma  mb  m, we have   , ra  and rb  , [in
2 2 2
this case the motion
Since there are no external forces, there is no external torques either].
This angular momentum of the system is conserved.
Since the system’s center-of-mass has been taken to be at rest at the
origin, the angular momentum consists only of the internal angular
momentum due to motion of the two particles about the center of mass.
This angular momentum is
    
L  ma ra  ra  mb rb  rb ,
(1.7)

68
PHY 301 CLASSICAL MECHANICS II


let’s re-write this in terms of rab
       
L  rab  ra  rab  rb  rab  rab  rab   ab
(1.8)
The two-body system begins to look like a single particle of mass μ and

coordinate rab .
The kinetic and potential energies of the system are
1  2 1  2 1 2   1 1  1  2
T ma ra  mb rb   rab     rab
2 2 2  ma mb  2 ,
U  U (rab )
(1.9)
The Lagrangian is
1  2
L rab  U (rab ) .
2
(1.10)
The Lagrangian is identical to that of a single particle system with mass
 
μ and coordinate rab . Since L is conserved, we know the motion is
 
restricted to the plane defined by rab and ab . Let this plane define a
spherical polar coordinate system (rab , ab , ab ), where ab , is the
azimuthal angle of the plane and ab is the polar angle of the position

vector rab relative to the z-axis in the plane. Rewriting L in this system
gives
L
1
2
 
 rab 2  rab 2 ab 2  rab 2 sin 2  abab  U (rab ) .
(1.11)
We may choose ab  0 without loss of generality. The angular
momentum vector points out of this plane (in the y direction) and the
motion remains in this plane at all time by conservation of angular
momentum.

3.1.2 Reduction to a one body problem

For simplicity and better understanding of the above equation of motion


we discuss the reduction of the two body problem to a mathematically
equivalent problem of a single particle moving in one direction. First we
reduce it to a one-body problem, and then we reduce the dimensionality.
Our original problem has six degrees of freedom, but because of the
symmetries in the problem, many of these can be simply separated and
solved for. As there are no external forces, we expect the center of mass
coordinate to be in uniform motion, and it allows us to use
  

R
 m r 
m1r1  m2 r2
 m m1  m2
(1.12)

69
PHY 301 CLASSICAL MECHANICS II

as three of our generalized coordinates. For the other three, we first use
the Cartesian components of the relative coordinate
  
r  r1  r2
(1.13)
although, we will soon change to spherical coordinates for this vector. In

terms of R and r the particle positions are:
  m    m 
r1  R  2 r r2  R  2 r
M M
(1.14)
where M=m1+m2.
The kinetic energy is
1  2 1  2
T m1 r1  m2 r2 ,
2 2
2 2
1   m2   1   m1    1 m1 m2  2
r   m1  m2 R 2 
1
T  m1  R  r   m2  R  r ,
2  M  2  M  2 2 M
1  1 
T  MR 2  r 2 ,
2 2
(1.15)

m1m2
where   is called the reduced mass. Thus the kinetic energy is
m1  m2
transformed to the form for two effective particles of mass M and μ
which is neither simpler nor more complicated than it was in the original
variables.
For the potential energy, however, the new variables are to be preferred,

U  r1  r2   U  r  is independent of R, whose three components are
  

therefore ignorable coordinates, and their conjugate momenta.


P   TR U   MR

cm i i
i

(1.16)
are conserved. This reduces half of the motion to triviality, leaving an
1  2
effective one-body problem with T  r and the given potential to
2

U (r ) . We have not yet made use of the fact that U only depends on the

magnitude of r . In fact, the above reduction applies to any two-body
system without external forces, as long as Newton's third Law holds.

Self Assessment Exercise A


1. Use momentum conservation to reduce the two-body problem to the
problem of one body motion in a central force field.

70
PHY 301 CLASSICAL MECHANICS II

3.2 Dynamics of an Isolated Two-Body Central-Force System

Now, let’s explore the dynamics using the Lagrangian. Conservation of



L already leads us to expect that one of the Euler-Lagrange equations
will be trivial. Explicitly, we have (dropping the ab subscripts now):
dU
r    r 2  r sin 2  2
dr
(1.17)
d
dt
 
r 2  r 2 sin cos 2
d
dt
 
r 2 sin 2   0
(1.18)
We make the general point that, when writing the Euler-Lagrange
equations for a multidimensional system, it is a good idea to start with
the time derivatives on the left side unexpanded until the rest has been
simplified because they are just the time derivatives of the canonical
momenta and some of them may end up being conserved if the right side
of the corresponding equation vanishes, either explicitly or by
appropriate choice of initial conditions. We see in the above case that
the ϕ equation of motion tells us lϕ = constant, which would have
become very unobvious if the derivative had been expanded into its
three terms.
  
For initial conditions, we take r and p to be in the plane ϕ= 0. That p
is in the plane ϕ=0 also implies ϕ= 0 initially. With these initial
conditions, the ϕ equation of motion implies that ϕ = 0 for all time. Thus,
l  r 2 sin 2   0 , for all time. The ϕ equation then tells us we
have l  r 2  cons tan t . The angular momentum vector has length

L  l and points perpendicular to the plane ϕ = 0 in which the motion

occurs: L is along the y-axis. The r equation of motion simplifies to
2
 dU l
r   .
dr r 3
(1.19)
The equation of motion is now that of a single particle in one dimension
with the effective
potential function
2
l
U eff (r )  U (r )   2 .
2r
(1.20)
We acquire a new ―centrifugal potential‖ that arises due to conservation
of angular momentum.

71
PHY 301 CLASSICAL MECHANICS II

It is a repulsive potential, reflecting the fact that, with lϴ constant, the


kinetic energy must increase as r−2 if r decreases then more energy is
needed to go to small radii. We may use the effective potential to write
an effective one-dimensional Lagrangian:
1 2 l
L1D  r   2  U (r ) .
2 2r
(1.21)
Note that the 1D Lagrangian is not obtained simply by rewriting the 3D
Lagrangian using ϕ = 0, ϕ2= 0 and μr2ϴ=lϴ one would have gotten the
wrong sign for the centrifugal term.
This difficulty occurs because the Lagrangian formalism assumes
independent variations of
the different coordinates. Instead, we must simply start with the
effective potential.
The effective total energy is
2
1 l
E  r 2   2  U (r )
2 2r
(1.22)
which is conserved because the effective potential is conservative. This
effective total energy
turns out to be equal to the true total energy because the ϴ kinetic
energy is included via the lϴ term.

3.2.1 Reduction to one dimension

In the problem under discussion, however, there is the additional



restriction that the potential depends only on the magnitude of r , that is,

on the distance between the two particles, and not on the direction of r .
Thus we now convert from cartesian to spherical coordinates (r; ϴ; ϕ),

for r . In terms of the cartesian coordinates (x; y; z)
1
r  (x 2  y 2  z 2 ) 2 ; x  r sin  cos ,
 
  cos1 z r ; y  r sin  sin  ,

  tan 1  y z ; z  r cos ,


 
(1.23)
Plugging into the kinetic energy is messy but eventually reduces to a
rather simple form

72
PHY 301 CLASSICAL MECHANICS II

T
1
2

 x1 2  y 2  z 2
1
 [( r sin  cos  r cos cos  r sin  sin  ) 2
2
 (r sin  sin   r cos sin   r sin  cos ) 2  (r cos  r sin  ) 2 ]


1
2

 r 2  r 2 2  r 2 sin 2  2 

(1.24)
Notice that in spherical coordinates T is a function of r and ϴ as well as
r, and , but it is not a function of ϕ, which is therefore an ignorable
coordinate and
L l
P    r 2 sin 2   cons tan t .
 
(1.25)
Note that r sin ϴ is the distance of the particle from the z-axis, so Pϕ is
just the z-component of the angular momentum, lz Of course all
  
component of the angular momentum L  l  r  p is conserved,
because in our effective one body problem there is no torque about the
origin. Thus l is a constant and the motion must remain in a plane
perpendicular to l and passing through the origin as a consequence of

the fact that r is perpendular to l . It simplifies things if we choose our

coordinates so that L is in the z-direction. Then    ,   0, 2
l  r  . The r equation of motion is then
2

2
l
r  r 2  dU  0  r   3  dU .
2
dr r dr
(1.26)
This is the one-dimensional motion of body in an effective potential
2
l
U eff (r )  U (r )  .
2r 2
(1.27)
Thus we have reduced a two-body three-dimensional problem to one
with single degree of freedom, without any addition of a centrifugal
2
l
barrier term to the potential.
2 r 2

Self Assessment Exercise B

1. Use total angular momentum conservation to show that one body in


a central force field moves in a plane. Write corresponding
Lagrangian in polar coordinates.

73
PHY 301 CLASSICAL MECHANICS II

2. Use angular momentum conservation to reduce the problem to an


analysis of a one-dimensional motion.
3. Give the expression for the centrifugal potential energy.

3.3 The formal Solution to The Equation Of Motion

We have obtained two equations of motion


U eff r , l ,
d
r 
dr
d
dt
 
r 2  0.
(1.28)
Let’s attempt to integrate these two equations. Integrating the r equation
most obviously yields as equation for r . But we already know what
equation will be, by energy conservation.

3.3.1 Integration of Equation of motion

We can simplify the problem even more by using the conservation law,
that of energy. Because the energy of the effective motion is a constant,
1 2
E r  U eff  cons tan t ,
2
(1.29)
we can immediately solve for
12
dr 2 
r     ( E  U eff (r )) .
dt  
(1.30)
This can be inverted and integrated over r, to give
dr
t  t0  
2( E  U eff (r )) / 
(1.31)
which is the inverse function of the solution to the radial motion
problem r(t). We can also find the orbit since
d  l dt
  2
dr dr / dt r dr
(1.32)
r
dr
  0  l  .
r0 r 2
2 ( E  U eff (r )
(1.33)
The sign ambiguity from the square root is only because r may be
increasing or decreasing, but time, and usually ϕ/L are always increasing.
Qualitative features of the motion are largely determined by the range
over which the argument of the square root is positive, as for other
values of r we would have imaginary velocities. Thus the motion is

74
PHY 301 CLASSICAL MECHANICS II

restricted to this allowed region. Unless L = 0 or the potential U(r) is


very strongly attractive for small r, the centrifugal barrier will dominate,
so U eff   and there must be a smallest radius rp > 0 for which
E  U eff .

3.3.2 A Differential Relation between r and ϴ

While we have formally eliminated t and obtained an integral


relationship between ϴ and r, the fact that the integral is not in general
analytic limits its usefulness. It may be more useful to have a differential,
rather than integral, relation between r and ϴ. We need to eliminate
d/dϴ from our original differential equations. The ϴ equation tells us
that
r 2 d l d
dt  d ,  2 .
l dt r d
(1.34)
Our equation of motion for r can then be re-written:
l d  l dr  l
2

   F (r )
r 2 d  r 2 d  r 2
d 1 dr  1 r 2
  2 F (r )
d  r 2 d  r l
d  d 1  1 r 2
   2 F (r )
d  d r  r l
d2 1 1 r 2
     2 F (r ).
d 2 r r l
(1.35)
We now have a differential equation with ϴ as the independent variable
and r as the dependent variable. The equation may be useful for
obtaining the shapes of the orbits, and frequently written via a change of
variables to u=1/r in the form
d 2u  1
 u   2 2 F  .
d 2
l u  u 
(1.36)
The constant of the motion, the energy, can be rewritten in terms of u
and ϴ alone (i.e, eliminating explicit time derivatives):

75
PHY 301 CLASSICAL MECHANICS II

2
1 l
E  r 2   2  U (r )
2 2 r
2 2 2
l  1 dr  l
   2    2  U (r )
2   r d  2r
l
2
 du  2  1
E    u   U  .
2

2  d   u
(1.37)

Self Assessment Exercise C


1. Use energy conservation to find formally an orbit of a body moving
in central field in coordinates r, ϴ.
2. Derive the relation between r and ϴ

3.3.3 Qualitative Dynamics

It is instructive to consider the shape of the effective potential energy


function and its implications for the motion of the system. The effective
potential consists of two terms. The first is the true central force
potential. The second is a ―centrifugal‖ term: it is a repulsive potential
arising from angular momentum conservation, which requires the kinetic
energy to increase as r is reduced. The relative sizes of the two terms
determine if or whether the effective potential is attractive or repulsive.
The shape of the effective potential and the total energy of the system
determine whether the orbits are unbounded, bounded, or bounded and
circular, and whether bounded orbits are periodic (closed). To be clear:
bounded and unbounded refers to whether there is an upper limit on r or
not; open and closed refer to whether the orbit repeats itself after some
period. All unbounded orbits are open, but not all bounded orbits are
closed.
The effective potential is:
l
U eff (r )  U (r )  .
2r 2
(1.38)

Consider different cases for the shape of U(r). Some of these are
illustrated in the Figure 1.0.
 Repulsive Potentials: If U(r) has no attractive regions (no regions
with positive slope), then both terms are repulsive at all r and all
orbits are unbounded and open. This occurs, for example, for the
Coulomb force between particles of like charge.
 Small r Behaviour: The small r behaviour of the effective
potential determines whether r is bounded below or whether there
are ―small r‖ bounded orbits with r bounded above.

76
PHY 301 CLASSICAL MECHANICS II

Figure 1.0: Effective potential for various true potentials.


The potentials are U(r) = rn or U(r) =logr with unity coefficient
except for the r−2 case. For all cases except r−2, the behaviour is
2
l
qualitatively independent of the factor  , so it has been set to 1.
2
2
−2 l
For the r potential, we show the three cases  = 0.5, 1, 1.5.
2
The locus of r values filled by example bound orbits are indicated
by dotted lines, by example orbits bounded below but unbounded
above by dashed lines, and by completely unbounded example
orbits by dash-dot lines, respectively.
 Large r Behavior: The large r behavior of the effective potential
determines whether r is bounded above at large r, which is a
necessary condition for bounded orbits.
 Circular orbits are obtained when there is a point in the effective
potential where the gradient (effective force) vanishes. For
potentials steeper than r−2, there is one unstable circular orbit. For
potentials shallower than r−2 and for log or positive power law
potentials, there is one stable circular orbit.
 Of course, if the true potential is a more complicated function of r
than simple power law or log functions, there may be additional
bound orbits and circular orbits.

77
PHY 301 CLASSICAL MECHANICS II

 Whether an orbit is periodic (that is, ―closed‖ the orbit repeats on


itself after some time) is a nontrivial problem. Clearly, all bound
orbits show periodic motion in r as a function of time. But
periodicity in space that is, the periodicity of r(ϴ) is a more
complicated problem. All circular orbits are periodic in space
because r is constant. More generally, periodicity of r(ϴ) requires
that the time periods for radial motion and angular motion be
commensurate that is, their ratio must be a rational fraction. It is
not obvious what the relation is between the form of the potential
and whether this condition is satisfied. Bertrand’s theorem tells
us that bound orbits are closed only if the potential in the vicinity
of the bound orbits follows r−1 or r2. Details on Bertrand’s
theorem can be found in Goldstein.
When orbits are bounded, the two turning points are called the apsides
or apsidal distances.
When orbits are not closed, the apsides precess in (r, ϴ) space. The
angle between two consecutive apsides is called the apsidal angle.
Precession of the apsides occurs whenever the conditions of Bertrand’s
theorem are not satisfied, including small perturbations of Bertrand’s
theorem potentials by non-Bertrand’s theorem terms.

Self Assessment Exercise D


1. Discuss briefly the shape of the effective potential energy function
and its implications for the motion of the system

4.0 Conclusion

Central force is defined as one that satisfies the strong form of Newton’s
third law. That is, given two particles a and b, the force exerted by
particle a on b is equal and opposite to that exerted by particle b on
particle a, and, moreover, the force depends only on the separation of the
two particles and points along the vector between the two particles.

Considering an isolated two-body system interacting via a conservative


central force. The dynamics of the two-particle system is equivalent to
that of a single particle with mass μ moving in the potential U(rab) and
likewise a two-body three-dimensional problem is equivalent to one
with single degree of freedom, without any addition of a centrifugal
2
l
barrier term to the potential. The shape of the effective potential
2 r 2
energy function and its implications for the motion of the system was
also considered for various true potential and it’s depicted in figure 1.0
above

78
PHY 301 CLASSICAL MECHANICS II

5.0 Summary

 Generic Central Forces.


The problem of two particles interacting via a strong-form third law
central force can be reduced to translational motion of the center-
of-mass system combined with one particle in a central force.
 
If the two particles have masses ma and mb and position ra and rb , then
the relative coordinate and reduced mass are
   ma mb
rab  ra  rb  .
ma  mb
 The original coordinates can be rewritten as
     
ra  rab rb  rab .
ma mb
 The lagrangian can be rewritten in the following ways
L
1  2
2
1 
 
rab  U rab    rab2  rab2 ab2  rab2 sin 2  abab2  U rab  .
2
 We obtain equations of motion (after eliminating the constant ф
coordinate)
dU l2
r    3 r 2  l  cons tan t .
dr r
 We define the effective potential
l2
U eff (r )  U (r )  .
2r 2
 The Lagrangian and energy can be rewritten in one dimensional
form
1 2 l2 1
L1D  r   2  U (r )  r 2  U eff (r ) ,
2 2r 2

1 2 l2 1
E r   2  U (r )  r 2  U eff (r ) ,
2 2r 2
l2
where is the repulsive ―centrifugal potential‖ The quantitative
2 r 2
behaviour of the system can be obtained by examining the shape
of the effective potential: where it is repulsive, attractive, where
its slope vanishes e.t.c The constancy of lθ gives us kepler’s
second law
dA 1 l
  cons tan t .
dt 2 
 The generic quadrature solution to the central force problem is given
by
1 2
r
2 l2 
t    dr  E  U (r )   2 2 
r (0)   r 

79
PHY 301 CLASSICAL MECHANICS II

l t dt 
  r (t )
   (0)   2
.
0

Elimination of t from the original differential relations also


allows us to obtain the quadrature solution for θ in terms of r.
1 2
dr   l2 
r
 (r )   (0)  l   2   E  U ( r )   
r ( 0 ) r (t ) 
2
 r 2 
.
 We may obtain a generic differential equation relating r and θ
d2 1 1 r 2
     2 F (r ) .
d 2  r  r l
1
Defining u  , this may be re-written
r
d 2u  1
 u   2 2 F  .
d 2
l u  u 
 The total energy of the system is constant and can be written as
l2  du   1
2
l2  1 dr  l2
E  2   U (r )     u   u 
2

2  r d  2r 2
2  d    u 

6.0 Tutor Marked Assignments (TMAs)

1. Consider a particle constrained to move on the surface described in


cylindrical coordinates by z  r 3 , subject to a constant

gravitational force F  mgeˆ z . Find the Lagrangian, two conserved
quantities, and reduce the problem to a one dimensional problem.
What is the condition for circular motion at constant r?
2. From the general expression for ϕ as an integral over r, applied to a

three dimensional symmetrical harmonic oscillator V (r )  12 kr 2 ,
integrate the equation, and show that the motion is an ellipse, with
the center of force at the center of the ellipse. Consider the three
complex quantities Qi  pi  i kmri and show that each has a very
simple equation of motion, as a consequence of which the nine
quantities Qi*Qk are conserved. Identify as many as possible of these
with previously known conserved quantities.
3. Show that if a particle under the influence of a central force has an
orbit which is a circle passing through the point of attraction, then
the force is a power law with F r 5 . Assuming the potential is
defined so that U    0, show that for this particular orbit E = 0,
find the period, and by expressing x , y and the speed as a function
of the angle measured from the center of the circle, and its
derivative, show that x , y and the speed all go to infinity as the
particle passes through the center of force.

80
PHY 301 CLASSICAL MECHANICS II

7.0 Further Reading and Other Resources

Hand, L. N. and Finch, J. D. (1998). Analytical Mechanics. Cambridge


University Press.

Thornton, S.T. and Marion, J. B. (2014.). Classical Dynamics of


Particles and Systems, Fifth Edition. Thomson Books/Cole

Goldstein, H., Poole, C., and Safko, J. (2000). Classical Mechanics,


Third Edition. Horoloma Book Jovanovich College Publisher.

Video Link:
https://www.youtube.com/playlist?list=PLtUBquogGkRukyAAWOmq2
LmQx8Yf9H_3E

81
PHY 301 CLASSICAL MECHANICS II

UNIT 2 KEPLER‖S PROBLEM

CONTENTS

1.0 Introduction
2.0 Objectives
3.0 Main Contents
3.1 Kepler’s Law
3.2 The Kepler Problem: The Special Case of Gravity
3.2.1 The Shape Of Solution Of The Kepler’s
Problem
3.2.2 Detailed Study of the Different Solutions
3.2.3 Summary of Quantities
3.2.4 Time Dependence of the Kepler Problem
Solutions
4.0 Conclusion
5.0 Summary
6.0 Tutor Marked Assignment
7.0 Further Reading and Other Resources

1.0 Introduction

The Kepler problem is named after Johannes Kepler, who proposed


Kepler's laws of planetary motion (which are part of classical mechanics
and solved the problem for the orbits of the planets) and investigated the
types of forces that would result in orbits obeying those laws (called
Kepler's inverse problem).
In classical mechanics, Kepler’s problem is a special case of the two-
body problem, in which the two bodies interact by a central force F that
varies in strength as the inverse square of the distance r between them.
The force may be either attractive or repulsive. The "problem" to be
solved is to find the position or speed of the two bodies over time given
their masses and initial positions and velocities. Using classical
mechanics, the solution can be expressed as a Kepler orbit using six
orbital elements.
2.0 Objectives
By the end of this unit, you will be able to:

 State the three Kepler’s Laws.


 Prove the three Keplers laws.
 Define an orbit.
 Derive and explain the conic equation of an orbit.
 Explain Keplerian orbit.
 Solve related problems.

82
PHY 301 CLASSICAL MECHANICS II

3.0 Main Contents

3.1 Kepler’s Law


r
dr
Practical integration of equation   0  l  is in
r0 r 2
2 ( E  U eff (r )
general a formidable task. The problem can be solved explicitly,
k
however, for the potential of the form U (r )   corresponding for
r
example to the gravitation field. Consider motion of a planet of mass mp
around the sun of mass ms. It was shown that this problem is equivalent
to an analysis of a motion of one body of reduced mass
1 1 1
  in central field U(r) around the center of mass
 m p ms
m p rp  ms rs
R ,
m p  ms
where rp and rs are the position of the planet and of the sun and
r  rp  rs . The corresponding effective Lagrangian,
1
L  (r 2  r 2 2 )  U (r )
2
Lagrange’s equations in polar coordinates are:
d  L  L d  L  L
   0. and    0,
dt  r  r dt    
Using the second equation above to prove kepler’s second law, note that
L does not depend on θ, therefore
d  L  L L
    0,  p   r 2  l
dt     
(2.1)
1
where lϴ is a constant. The radius vector r sweeps out an area dA  r 2
2
in time dt. The rate at which the radius vector sweeps out area
dA 1 2 
is  r  Comparing this rate with the momentum p we proved
dt 2
dA 1 2  1
that  r  cons tan t
dt 2 2
(2.2)
which is Kepler’s second law which state that the radius vector drawn
from the sun to a planet describes equal areas in equal times.
Using first Lagrange’s equation we can prove Kepler’s first law which
states that each planet moves in elliptical orbit with the sun at one focus
d  L  L
 
dt  r  r
 
 0   r  r 2  
U (r )
r
k
 2 .
r
(2.3)
Proof:

83
PHY 301 CLASSICAL MECHANICS II

l u 2 
r 2  l     dt  d ,
 lu 2
(2.4)
1
since u  .
r
d l u 2 d
The equation above can rewritten as follows:  , also, r can
dt  d
be re-written as
dr 1 du  du
 2 
dt u dt l d
d  dr  l u 2 d  1 du 
therefore     on substituting for r and  in the
dt  dt   d   d 
equation, we have

k
 
2 2
 k l u 2 d 2 u l u 3 d 2u
 r  r   2  
2
   ku 2
 u  2 .
r  d 2
 d 2
l
(2.5)
Thus, equation for the planet motion for linear oscillator in a constant
field. Its solution is a sum of a general solution b cos(  0 ) and a
k
particular solution . i.e
l
k
u 2
 b cos(  0 ) .
l
(2.6)
Or, in polar coordinates, we have

r
1  p cos(   0 )
(2.7)
2 2
l qEl
where    and p  1  2  .
k k 
(2.8)
This is an equation for conic sections which describes
 p  1 hyperbola

 p  0 parabola
 p  1 ellipse.

(2.9)
Motion of the planet is bounded to the sun and therefore corresponds to
the case p<1. Futhermore, since ms is far less than mp (ms=333,500 x
mearth and mearth=5,977x1024kg), we have
1 1 1 1
  
 m p ms m p

84
PHY 301 CLASSICAL MECHANICS II

m p rp  ms rs mp ms
and R   rp  rs  rs
m p  ms m p  ms m p  ms
i.e. centre of mass approximately coincides with the position of the sun
and r is approximately distance from the sun to the planet. Therefore, we
have proved Kepler’s first law which states that each planet moves in
elliptical orbit with the sun at one focus.

dA 1 2
Using the second Kepler,s law  r  , and the expression for the
dt 2
1
angular momentum p  r 2 to write dA  dt Integrating this
2
expression over the whole area of the ellipse, we obtained
1
A  ab  T , where T is the period of planetary motion and πab is the
2
area of the ellipse. Using the following relation a   1b2 to obtain
1 3 1 l 2 1 2
 2 a 2  T  a3  T .
2 (2) 2
l2 k
Using the definition   to finally get a3  2 T 2 which is the
k 4 
Kepler’s third law which states that the square of the period of
revolution about the sun is proportional to the cube of the major axis of
the orbit.

Self Assessment Exercise A


1. Use ϴ-component of Lagrangian equation to prove Kepler’s second
law.
2. Use r-component of Lagrange’s equation to prove Kepler’s first law.
3. Use the second Kepler’s law and the expression for the angular
momentum to prove Kepler’s third Law
4. Derive the equation of a conic section and describe the shape for p >
0, p = 0 and p < 0

3.2 The Kepler Problem: The Special Case of Gravity

We now specialize to gravity, which allows us to fix the form of the


central-force potential energy function. We solve the equation of motion
and study the various solutions with k=GμM.

3.2.1 The Shape of Solution of The Kepler’s Problem

The General Solution


In our generic study of central forces, we obtained the differential
relation in equation 1.36

85
PHY 301 CLASSICAL MECHANICS II

d2 1 1 r 2
     F (r ) .
d 2  r  r l
mm
For the gravitational force, we have, F (r )  G a 2 b ,
r
(2.10)
so the above reduces to
d2 1 1 G 2 (ma  mb )
     .
d 2  r  r l
2

(2.11)
1
Re-writing using u  , we have
r
d 2u G 2 M
u   ,
d 2 l
2

(2.12)
ma mb
where M  ma  mb  .
This is now a simple harmonic oscillator equation with a constant
driving force. The solution is the sum of the generic solution to the
homogeneous equation and a particular solution to the inhomogeneous
equation:
G 2 M
u ( )  A cos(  0 )  2
.
l
(2.13)
We can relate the coefficient A in the solution to the constants of the
motion, the energy and angular momentum by using equation 1.37
l  du  
2 2
1
E    u   U  ,
2

2   d   u

l  
2
G 2 M  G 2 M 
2

   A 2 sin 2 (   0 )  A 2 cos2 (   0 )  2 A cos(   0 )   


2  2 2  

l  l  
 G 2 M 
 GM  A cos(   0 )  2
,

 l 
G2 2M 2 G2 2M 2
2
l A 2
  AGM cos(   0 )   GMA cos(   )  ,
2 2 0 2
2l l
l A 2 G 2  2 M 2
2

  ,
2 2l
2

l  
2
2
 G 2 M 
   A 2   

.
2   l
2
 
 
(2.14)

86
PHY 301 CLASSICAL MECHANICS II

Let’s us write the orbit in terms of r instead of u and also drop the offset
phase θ0,
p
 1   cos  p  r  r cos
r
(2.15)
2
l
where p  and   pA .
G 2 M
If we write in Cartesian coordinates, with x  r cos and y  r sin , we
have
p  x 2  y 2  x   p  x   x 2  y 2
2

(2.16)
1   2 x2  2px  y2  p2  0 .
(2.17)
In terms of p and ε, the total energy is
l  2 1 
2
E    2  2 ,
2  p p 
(2.18)
G 2 3M 2 2
E 2
  1
2l
GM 2

2p

 1 . 
(2.19)
Let’s re-write in a more obvious form: complete the square on x to
obtain
1   2  x  p 2   y 2  p 2
2

 1   1 
x  xc 2  y 2 1,
a2 b2
(2.20)
which is the equation for a conic section with
p p p
xc  , a , b f  xc  a  0 and 2 xc
1  1   1   2 
2 2

where f denotes the x coordinates of the foci of the conic section. Recall
that the center of mass of the system is at the origin, so one of the foci
coincides with the center of mass. The  sign is picked depending on
the sign of 1   2 to ensure that b is real. The turning points of the
motion are given by the maximum and minimum values of r. Our polar
form for the orbit, equation 2.15 provides the easiest means to obtain
these: they are cos  1 they are therefore
p p
r1  r2 
1  1 
(2.21)

87
PHY 301 CLASSICAL MECHANICS II

p p
x1  x2 
1  1 
(2.22)
y1  0 y2  0
Where x1, 2  r1, 2 cos so x pick up a sign and y1, 2  r1, 2 sin , so y vanishes
in both cases. The energy is now
GM 2 GM
E (  1)   .
2p 2a
(2.23)
The sign and magnitude of the energy thus scales inversely as the semi
major axis a.
So what we have is the equation of a conic section. Being a circle,
ellipse, or hyperbola depending on the sign of 1   2 . From the
qualitative discussion of central force orbits, E=0 is the dividing line
between bound and unbound orbits. The implication then is that bound
orbits with E<0 have positive a and ε2< 1 while unbound orbits with
E>0 have negative a and ε2>1. The dividing case is E=0, a   , ε =1.
The conic section formula is undefined there, but if we go back before
we completed the square, we see that ε2=1 causes x2 term to vanish
leaving us with the equation n of a parabola (x a quadratic function of y).

Self Assessment Exercise B


1. Derive the equation for a conic section of an orbit and discuss he
nature of the curve for every possible value of e.

3.2.2 Detailed Study of the Different Solutions

Let’s study these various solutions in some details. First, let’s obtain a
dimensionless parameterization of the solutions. The shape of the
effective potential is set by lθ and μ. The effective potential is minimized
when the effective force vanishes. Equation 1.19 we have
GM
2 2
l l
 3 r  p.
r2 r G 2 M
(2.24)

The value of the effective potential at this point, which gives the
minimum physically allowed value of the total energy is
GM
2 2
l l
E min  U eff (r  p )    ,
2 p 2 p 2 p 2
1 GM
   E scale ,
2 p
(2.25)

88
PHY 301 CLASSICAL MECHANICS II

where we have defined a scale energy that is the absolute value of the
minimum energy. Referring back to our equation for E in terms of ε and
p. Equation 2.19, we see that
E  EScale ( 2  1)
(2.26)
With Emin and Escale in hand, let’s consider the various cases. Examples
are illustrated in figure 2.1 below
 E / Escale  1 : not physically allowed.
 E / Escale  1 : Equation 2.26 tells us that E=-Escale corresponds to ε2
= 0. Since the eccentricity vanishes. The solution from Equation
2.15 is p = r for all θ; i.e. the orbit is a circle. This is as one
would expect from the effective potential that if the solution is at
minimum of the effective potential then, there is no radial force.
The conic section solution is elliptical (because ε2 < 1) and the
semimajor axes are equal a=b=p as one would expect for a circle.
 −1 < E/Escale < 0: Because the energy remains negative, Equation
2.26 implies that 0 < ε2 < 1 and the conic section solution is
an ellipse. As the energy increases, the eccentricity of the ellipse
increases. Remember that the center of mass coincides with one
of the foci of the ellipse. The center of the ellipse xc and the
second focus 2xc move off to   as   1 .
 E/Escale = 0: For this solution, Equation 2.26 tells us that ε2=1. Our
derivation of the conic section form fails here because the
coefficient of the x2 term vanishes, but we can return to the
Cartesian form of the solution to find
2 px  y 2  p 2  0
p y2
x 
2 2p
(2.27)
This is a parabola whose vertex is at p/2, whose focus is at the origin,
and whose directrix is at p. Recall that the directrix is the line
perpendicular to the axis of the parabola such that the parabola consists
of the set of points equidistant from the focus and the directrix. The
system is just barely not bound, with the radial kinetic energy and
velocity approaching zero as r   , because the total energy vanishes
and the effective potential energy approaches zero as r   .

 E/Escale > 0: For this solution, Equation 2.26 gives ε2 > 1. The
conic section is a hyperbola. A hyperbola has two branches.
Because the polar form of the solution, Equation 2.15 implies a
one-to-one relationship between r and θ, only one branch of the
hyperbola can be a valid solution. Intuitively, based on
continuously transforming the eccentricity, we expect this to be
the left branch. We can see this explicitly as follows. The left and

89
PHY 301 CLASSICAL MECHANICS II

right branches are distinguished by the fact that the former has
regions with x < 0, while the latter does not. In order to have
negative values of x, θ must be allowed to go outside the range (-
π/2, +π/2). A restriction on θ is placed by the requirement that r
1
be positive, which translates to the requirement cos   . Since

ε2 > 1 (and ε is taken to be positive always), this defines a
maximum value of |θ| that is between π/2 and π. Hence, x is
allowed to be negative, and so the left branch solution is the
appropriate one. For the hyperbolic solution, there is only one
turning point, which is the x1 turning point. Let’s consider the
evolution of the solution with ε. One focus of the hyperbola
always remains at the origin. The ―center‖ of the hyperbola, xc,
starts out at +1 and moves in toward the origin as ε gets larger,
with xc  0 in the limit    .Thus, the hyperbola continuously
transforms from a parabolic-like orbit to a straight vertical line,
with the turning point x1 moving closer to the origin as ε
increases. These solutions are definitely not bound. They in fact
have excess kinetic energy so that, as r   , the radial kinetic
energy (and hence radial velocity) remains non zero.

A Note on Repulsive Potentials.


While we have so far considered only attractive potentials, it is
straightforward to translate the above solution to the case of repulsive
potentials. We will see that, for a given energy E, we obtain the same
hyperbola as for the attractive potential, but we must choose the right
branch, not the left branch.

Figure 2.1: Example Keplerian orbits. The left and right figures have
identical orbits; only the axis range is different. All these orbits have
2
l
and p = 1, so they have the same angular momentum (same
2
centrifugal barrier) but different total energies. The scale factor for the
energy

90
PHY 301 CLASSICAL MECHANICS II

2
l
is therefore also 1, so Escale = 1 and the various orbits have energy
2 p
E = ε2 − 1. The legend
is, in order of increasing eccentricity: ε = 0 (red), ε = 0.25 (green), ε =
0.5 (blue), ε = 1 (cyan),
ε = 2 (magenta), ε = 4 (yellow). The center of each orbit (xc) is shown by
the diamond of the same color, and the second focus by the squares. The
first focus of all orbits is at the origin. The
second branch of the hyperbolic orbits is not shown.

The repulsive potential solution can be obtained from the attractive


potential solution by simply taking GμM < 0 and making use of the
physical fact that only E > 0 is allowed for a repulsive potential. Let’s
go through the derivation with these changes. First, the solution for u
becomes
G 2 M
u    A cos  2
.
l
(2.28)
Since u  0 is required, the solution must have A > 0 and is only valid
for some range of θ. Keeping our original definition of p (which now
implies p < 0), our polar form of the solution is
p
 1   cos .
r
(2.29)
Since p < 0 and A > 0, we have ε= pA < 0 also. Since p < 0, the
solution is valid only when the right side is less than zero. (Originally,
our requirement was that the right side be greater than zero because both
p and r were positive.) The region of validity is given
1 1
by cos    . For there to be any region of validity, we must have
 
|ε| > 1, which implies that only hyperbolic solutions will be allowed.
Furthermore, the region of validity has cos θ > 0, so we must use the
right branch of the hyperbolic solution we obtained. The conversion
from polar to Cartesian coordinates goes as before, as does the
completion of the square. So the hyperbolic solution is still valid, as are
the formulae for xc, a, b, and f. xc has the same sign as in the attractive
hyperbolic case (since the sign flips in ε and p cancel each other), a and
b change sign (a > 0, b < 0 for all repulsive solutions). Because xc does
not change sign, the foci are in the same place as the attractive
hyperbolic case. The sign flips in a and b do not affect the shape of the
hyperbola since only a2 and b2 enter in the conic section formula. Thus,
we have the exact same hyperbola as in the attractive case, except that
the restriction on θ implies that we must now take the right branch, not
the left branch. This also means that the turning point is now x2, not x1.

91
PHY 301 CLASSICAL MECHANICS II

The energy expressions, Equations 2.19 and 2.23, hold without change.
The starting point for the energy equation, Equation 2.18 is insensitive
to the sign of ε and p. Equation 2.19 does not change meaning that
because the sign flips in GμM and p cancel each other, so |ε| > 1 still
gives E > 0 for all repulsive potential solutions. Equation 2.23 also
keeps its same sign because both GμM and a change sign.
Intuitively, the change from the left branch to the right branch reflects
the fact that an attractive potential turns the trajectory inward toward the
center of force while a repulsive potential turns it outward.

3.2.3 Summary of Quantities

The various quantities involved in Keplerian orbits are summarized in


Table below.
Quantity Symbol Formula(e) Sign Significance
Angular
momentum

 r 2 ≥ 0 Centrifugal
= 0 gives trivial potential
orbit (brings in effect
of ϴ motion)
Scale Escale 1 GM >0 Scale energy
energy 2 p = Emin for
attractive
potential
Scale p 2
l > 0 for Sets scale of
radius G 2 M attractive pot orbit
< 0 for
repulsive pot
eccentricity ε 1
E ≥ 0 attractive Sets shape of
E
scale
pot conic section,
E
 1 related to ratio
E
scale
< -1 repulsive of energy to
pot. scale energy
Orbit xc p = 0 circular

center 1  2
< 0 elliptical
 a > 0 hyperbolic
Semimajor a p > 0 circle/ Distance from
axis 1  2
elliptical xc to vertices
< 0 hyperbolic along major
attractive axis
Semimajor b p > 0 attractive Distance from
axis 1  2 pot. xc to vertices
< 0 repulsive along major
Pot axis
Turning x1 p >0 Turning points
points 1  2
of motion
<0 relative to CM
92
PHY 301 CLASSICAL MECHANICS II

x2 p circle/elliptical= focus 1

1  2 apsides for
> 0 hyperbolic circ./ellip.
orbits
For hyperbolic orbits, x1 is the turning point for attractive potentials, x2
the turning point for repulsive potentials

Table 2.1: Parameters for Keplerian orbits.

3.2.4 Time Dependence of the Kepler Problem Solutions

So far we have only found the orbit solutions as functions r(θ). This of
course describes much of the dynamics of the problem. But one does
indeed frequently want the orbit as a function of time, so we obtain that
result here.

Period of Elliptical Orbits


We can quickly obtain the period of elliptical orbits by using
Kepler’s second law, Kepler’s second law tells us that
dA l
 .
dt 2
(2.30)
The area of the ellipse is A = π a b, so the period is the time required to
sweep out the area of the ellipse,
A ab
T  .
dA l
dt 2
(2.31)
Let’s write this in terms of the parameters of the orbit p and ε to obtain
Kepler’s third law:
p2 1
T  2
1   
2 32
 GMp
a3
 2
GM
(2.32)
The period depends only on a. Of course, a encodes information about
the total energy E and the angular momentum lθ. The implication of
Kepler’s third law for the solar system is that all orbits should lie on a
single T 2 a 3 curve because M, dominated by the sun, is almost the
same for all planets.

Self Assessment Exercise C

93
PHY 301 CLASSICAL MECHANICS II

1. Prove that the scale energy is absolute value of minimum energy


and that E  EScale ( 2  1)
2. Prove that the Kepler’s law asserts that the period T of the motion
and the main radius a of the ellipse satisfy a relation T  Ca 3 / 2 ,
where C is a constant independent of the initial data.

4.0 Conclusion

Kepler tells us not only that the orbit is an ellipse, but also that the sun is
at one focus. To verify that, note the other focus of an ellipse is
symmetrically located, at (−2εa, 0), and work out the sum of the
distances of any point on the ellipse from the two foci. This will verify
that d+r = 2a is a constant, showing that the orbit is indeed an ellipse
with the sun at one focus.

5.0 Summary

 The Kepler’s problem


1
When we specialize to U (r )   , we may obtain more specific
r
results. The differential equation relating r and θ or u and θ is
d 2  1  1 G 2 M d 2u G 2 M
     u  .
d 2  r  r l2 d 2 l2
 The generic solution may be written
G 2 M l2
u ( )  A cos   0   p  r  r cos p 
l2 G 2 M
  pA .
We specify two initial conditions (neglecting θ0): and the total
energy E. The constant A and ε are related to the total energy by
l2  2 G 2 M  GM 2
E
2 
A  2
l 

2p
  1  Escale  2  1 .

 The energy may also be written as


GM
E
2a
 The assorted orbital parameters are summarized in table 2.1
 Kepler’s third law, which tells us that the period of elliptical
orbits, is obtained from Kepler’s second law and the area of an
elliptical orbit, giving as
a3
T  2 .
GM
 The full time dependence of elliptical orbits can be obtained
through use of the eccentric anomaly, ε, which is defined
implicitly in terms of the true anomaly, θ. One begins with

94
PHY 301 CLASSICAL MECHANICS II

r ( ) cos  x  xc  a cos
and one obtains
r ( )  a(1   cos ) t ( )  GMa 3    sin  
x( )  a(cos   ) y( )  a 1   2 sin  .
 An analogous parametization can be done for parabolic orbits,
though the geometrical interpretation of the eccentric anomaly is
no longer valid. Beginning with

r ( )  a( cosh  1)
t ( )  GMa 3  sin    
x( )  a(  cosh ) y( )  a 1   2 sin 
where the upper signs are for an attractive orbit and the lower
signs for a repulsive one.

6.0 Tutor Marked Assignments (TMAs)

1. Use ϴ-component of Lagrangian equation to prove Kepler’s second


law.
2. Use r-component of Lagrange’s equation to prove Kepler’s first law.
3. Use the second Kepler’s law and the expression for the angular
momentum to prove Kepler’s third Law.
4. Derive the equation of a conic section and describe the shape for p >
0, p = 0 and p < 0.
5. Derive the equation for a conic section of an orbit and discuss the
nature of the curve for every possible value of e.
6. Prove that the scale energy is absolute value of minimum energy
and that E  EScale ( 2  1) .
7. Prove that the Kepler’s law asserts that the period T of the motion
and the main radius a of the ellipse satisfy a relation T  Ca 3 / 2
where C is a constant independent of the initial data.
7.0 Further Reading and Other Resources
Hand, L. N. and Finch, J. D. (1998). Analytical Mechanics. Cambridge
University Press.
Thornton, S.T. and Marion, J. B. (2014.). Classical Dynamics of
Particles and Systems, Fifth Edition. Thomson Books/Cole
Goldstein, H., Poole, C., and Safko, J. (2000). Classical Mechanics,
Third Edition. Horoloma Book Jovanovich College Publisher.

Video Link:
https://www.youtube.com/playlist?list=PLtUBquogGkRukyAAWOmq2
LmQx8Yf9H_3E

95
PHY 301 CLASSICAL MECHANICS II

UNIT 3 SCATTERING CROSS SECTIONS

CONTENTS

1.0 Introduction
2.0 Objectives
3.0 Main Contents
3.1 Setting up the Problem
3.1.1 Initial Conditions
3.1.2 Scattering Angle
3.2 The Generic Cross Section
3.2.1 Incident Beam
3.2.2 Differential Cross Section
3.2.3 Total Cross Section
3.2.4 Calculating b(θs)
1
3.3 Potential
r
3.3.1 Finding b(ϴs)
3.3.2 Calculating the Differential Cross Section
3.3.3 The Total Cross Section
4.0 Conclusion
5.0 Summary
6.0 Tutor Marked Assignment
7.0 Further Reading and Other Resources
1.0 Introduction
Now that we have studied central force motion, obtaining qualitative
results for arbitrary potentials and specific results for 1/r potentials, we
have information about the dynamics of collisional interactions. We can
use this to develop the concept of scattering cross section, which
intimately uses the kinematics of collisions and the orbit information
from central force motion. The archetypal scattering problem, we will
consider is one involving a particle incident on a force center that
scatters the incident particle via a conservative central force with
potential U(r).
We have demonstrated that, a two-particle system interacting via a
conservative central force is equivalent to such a system when
considered in its center-of-mass frame.
2.0 Objectives
By the end this unit, you will be able to:
 State the expression for energy of a system of particle incident on
a force center subject to a scattering potential.
 Calculate the differential cross section.
 Calculate the total cross section.

96
PHY 301 CLASSICAL MECHANICS II

3.0 Main Contents

3.1 Setting up the Problem

3.1.1 Initial Conditions

As noted, we consider a particle incident on a force center subject to a


scattering potential. It is assumed that the particle is asymptotically free,
having enough energy to be in an unbound orbit. In order to have
unbound orbits, the potential energy must go to zero at large r. The
energy of the system is therefore
1
E  v 2 ,
2
(3.1)
where v is the asymptotic particle speed as r   . We will use v to
parametize the initial energy of the system, v only specifies an energy.
We must also specify the geometry. The only free parameter left is lθ,
which we can see specifies the geometry as follows. Since the system is
unbound, the trajectory must become straight lines at large radii,
reflecting the incoming and outgoing velocity vectors. For example, for
a 1 r potential, we know that the ingoing and outgoing velocity vectors
define the asymptotes of the hyperbolic orbit. We can calculate the
distance between the scattering center and these straight lines where
they come closest to the scattering center. This distance is defined to be
the impact parameter, usually associated with the symbol b. This is
displayed in Figure 3.1 below. The small line from the center to the
dotted line is the impact parameter. The impact parameter is related to lθ.
  
We know that l  r  r . At r  , we know that r points in the

direction along the asymptote and r points to the origin, so the angle
 
between r and r which we call  is subtended by the impact parameter

b. r is shown approximately in Figure 3.1 by the line that extend from
the center to the -10 point. So,
l  r v sin   vb
(3.2)
lθ specifies the geometry through b, which fixes the position of the
asymptote (v ) relative to the scattering center.

3.1.2 Scattering Angle

In Figure 3.1, the scattering angle, θs, is the angle between the incoming
and outgoing velocity vectors. θs is determined by E and lθ or
equivalently v and b, and the form of the potential function.

97
PHY 301 CLASSICAL MECHANICS II

Figure 3.1: Scattering impact parameter illustration


(where  * is the Scattering angle herein represented
as θs)

Self Assessment Exercise A


1. Derive the expression of the angular momentum of the particle
incident on a force center subject to a scattering potential in terms of
μ, E and b

3.2 The Generic Cross Section

3.2.1 Incident Beam

Now that we have made our definitions, let us bring in the concept of
differential cross section. Suppose we have a beam of incoming

particles, all with same velocity v. . Let the flux of particles F be the
number of particles passing through a unit area in a unit time. If the
beam has particle number density n, then the flux is
F  nv
(3.3)
Assume that the beam has a circular cross section and the axis of the
beam points directly at the scattering center. The incoming particles will
thus have a range of impact parameter values, ranging from b = 0 (along
the beam axis) to b = bmax (at the outer edge of the beam). The beam
radius is bmax and its cross-sectional area is A  bmax
2
.

98
PHY 301 CLASSICAL MECHANICS II

3.2.2 Differential Cross Section

The incident particles in the beam will be scattered into a range of


angles depending on their input impact parameters (and the beam
d
velocity). We define the differential scattering cross section, ( s , s ),
d
via the probability of an incident particle being scattered into the solid
angle dΩ in the direction ( s , s ) where θs is the polar angle measured
from the beam axis and s is the azimuthal angle around the beam axis.
If dN (s , s ) is the number of particles per unit time scattered into the
solid angle d at ( s , s ) , we define the differential cross section via the
relation
1 d dN ( s ,s )
( s ,s )d  .
A d FA
(3.4)
Let us explain the above. On the right hand side of the first line, the
denominator is the number of particles per unit time incident on the
target from a beam of flux F and cross-sectional area A. Since the
numerator is the number of particles per unit time that scatter into d at
( s , s ) , the right hand side is thus the fraction of particles that scatter
into d at ( s , s ) ; it is a probability. We include additional factors so
d
that is defined only by the scattering force, not by parameters of the
d
experiment. We have 1/F on the right hand side but none on the left side
because including the 1/F makes the ratio dN/F independent of F: if F
goes up, dN goes up proportionally. However, we include 1/A on the left
hand side to cancel the 1/A on the right side because dN (s , s ) may not
scale with A: if one adds cross-sectional area at a radius from which
particles do not scatter into the particular solid angle d at ( s , s ) ,
dN (s , s ) will not increase when A increases. Hence the different
treatment of F and A. Solving for the differential cross section gives
d 1 dN ( s ,s )
( s ,s )  .
d F d
(3.5)
d
The beam area A has dropped out, has units of area per steradian;
d
hence the name cross section. If we assume central force scattering, the
problem is azimuthally symmetric about the beam axis and we may
integrate over s so that dΩ = 2πsin θsdθs. Furthermore, we know for
central force scattering that there is a one-to-one correspondence
between b and θs for a given v . Therefore, the particles scattering into
the interval dθs in polar angle come from some range db in impact
parameter. Azimuthal symmetry of the problem allows us to integrate

99
PHY 301 CLASSICAL MECHANICS II

over azimuthal angle in the beam also. We may relate db and dθs by
requiring conservation of particle number:
d
F 2bdb   F 2 sin s d s .
d
(3.6)
A negative sign has been inserted under the assumption that the potential
decreases in strength monotonically with radius: if you increase the
impact parameter a little bit, the scattering angle should decrease, so a
positive db implies a negative dθs. Re-writing, we have
d b db
 .
d sin s d s
(3.7)
That is, if we know the function b(s , v ) , then, we can determine the
distribution of particles in scattering angle given a uniform incoming
beam.

3.2.3 Total Cross Section

d
Once one has calculated , it is formally a straightforward thing to
d
calculate the total cross section:
 
d d
   d  2  d s sin s  2  dbb .
d 0
d 0

(3.8)
As one would expect, the total cross section is related to the probability
that an incoming particle will be scattered to any angle. It can be viewed
as a total ―effective area‖ of the scattering center; the number of
particles in the incoming beam that will be scattered is the same as if
every particle within the central σ of the beam were scattered and all
others left untouched.

3.2.4 Calculating b(θs)

For a generic central potential, we can obtain a formula relating b and θs


by returning to the integral relations that define the orbit. Recall
Equation 1.34:
1 2
dr   l 
r 2
 (r )   (0)  l  2 2 ( E  U (r )  2 
r ( 0) r 
r 
(3.9)
θ is not θs, but it is related to it. Referring back to our scattering picture
above, the scattering angle θs is the complement of the angle between
the incoming and outgoing asymptotes. The angle between the
asymptotes is related to the orbit angle as r   . For repulsive
scattering, the angle between the asymptotes is just  out   in , where

100
PHY 301 CLASSICAL MECHANICS II

these are the asymptotic orbit angles for the incoming and outgoing
particles. For attractive scattering, since ϴ is measured from the +x-axis,
the angle between the asymptotes is 2   out  in . So the scattering
angle is
 s   2   out   in 
(3.10)
where – (negative) goes with the attractive potential. Next,  out   in is
twice the angle between θout or θin and θ=0. θ=0 is obtained when
r=rmin, the turning point. So we have
 2 1 2 
dr   l 

 
 s     2 l  2 
2 ( E  U (r )  2  
 rmin r (t )   r  
 
 
bdr   U (r ) b 2 
1 2


    2  2 1   2 .
   E r  
rmin r
 
(3.11)

Self Assessment Exercise B


d b db
1. Given the differential cross section  Calculate the
d sin s d s
total cross section and the b(ϴ).

1
3.3 Potential
r
1
For the potential, we can find the differential cross section
r
explicitly because we have explicit relationships between ϴ and r

3.3.1 Finding b(ϴs)

We demonstrated earlier that the azimuthal angle ϴ of the orbit relative


1
to the center of mass is limited to be cos  this gives us ϴout and ϴin

 1
 out ,in   arccos  
 
(3.12)
where out  in and the choice of sign for ϴin depends on initial
conditions. So we have

101
PHY 301 CLASSICAL MECHANICS II

 1
 out  in  2 arccos  
 
 1
 2 arccos  
  
 
(3.13)
where the – (negative) is for an attractive potential. So we have
  1  1 1 
 s     2 arccos       2 arccos  sin s
      2
(3.14)
where the two different potentials have yielded the same result. With
some works, we can write b in terms of ε. Starting with equation 2.19,
we have
2
l
this can be re-written E ( 2  1)
2 p 2

p 
 ( 2  1)
l 2E .

(3.15)
l 
Then,   2 1
GM 2E
GM
We have v b   2 1 .
2E / 
(3.16)
GM
that is b  2 1 .
2E
(3.17)
So we may now find b(ϴs):
GM  GM 
b csc2 s  1  cot s ,
2E 2 2E 2
(3.18)
s 
where we take positive square root because 0   .
2 2

3.3.2 Calculating the Differential Cross Section

db
We will need to calculate the differential cross section, so, taking
d s
the derivative:
db 1 GM  b  
 csc2 s  csc s sec s .
d s 2 2E 2 2 2 2
(3.19)

102
PHY 301 CLASSICAL MECHANICS II

Finally, using our formula for the differential cross section:


d 2 s 2 s  GM 
2
b db b2 1
  csc sec   .
d sin s d s 4 2 2  4 E  sin 4  s
2
(3.20)
Note that the result is independent of whether the scattering is attractive
or repulsive. This result is the well known Rutherford scattering formula,
first observed in the scattering of α particles (4He nuclei) from gold
nuclei in a foil. The observation that the scattering obeyed the simple
1
expectation of a potential was an important piece of evidence that the
r
atom consists of a charged nucleus surrounded mostly by empty space,
as opposed to a ―plum pudding‖ type model with electrons and protons
intermixed uniformly in the atom.

3.3.3 The Total Cross Section

If one tries to calculate the total cross section from the Rutherford
formula, one will end
1
up with an infinite result. This is because, for a potential, the
r
probability of scattering does not decrease sufficiently quickly with
increasing b because the effective area of the scattering center is infinite.
If the potential is made to converge more quickly (e.g., by multiplying
be an exponential decay), then a finite total cross section is obtained.

Self Assessment Exercise C


1. Derive the b(ϴ), the differential cross section and total cross section
1
for the potential
r

4.0 Conclusion

We have discussed the 1/r potential in terms of Newtonian gravity but of


course it is equally applicable to Coulomb’s law of electrostatic forces.

5.0 Summary

 For central-force scattering problems, we generally considered


unbound central-force orbits, rather than parametizing in terms of
energy E and angular momentum lϴ, we use the velocity v and
the impact parameter b (distance of closest approach to scattering
center). These parameter are related by
1 2
E v l  vb ,
2

103
PHY 301 CLASSICAL MECHANICS II

 We usually phrase scattering in terms of an incoming beam of


particles of number flux
F  nv .
 The differential scattering cross section gives us the area of the
beam that will be scattered into a solid angle dΩ:
d 1 dN
 .
d F d
 The differential cross section can be found if the relationship
between the input impact parameter and the scattering angle ϴs
(the angle between the incoming and outgoing trajectories) is
known
d b db
 .
d sin s d s
 The total scattering cross section, which gives us the effective area
of the scattering center, is
 
d d
   d  2  d s sin s  2  dbb .
d 0
d 0

 The scattering angle is calculated from the potential function via


 bdr   U (r ) b 2  
1 2

 s     2  2 1   2
 r  E r  
 
k
 For a potential U (r )   , the relationship between impact
r
parameter and scattering angle and the differential cross section
are given by
 d  k 
2
k 1
b  cot s   4 
.
2E 2 d  4 E  sin s
2

The relative sign between b and ϴs is chosen based on whether the


potential is attractive or repulsive (E > 0 always for unbound
orbits)

6.0 Tutor Marked Assignments (TMAs)

1. Consider a spherical droplet of water in the sunlight. A ray of


light with impact parameter b is refracted, so by Snell's Law n sin
β = sinα. It is then internally reflected once and refracted again on
the way out.

104
PHY 301 CLASSICAL MECHANICS II

a) Express the scattering angle ϴ in terms of α and β.


b) Find the scattering cross section dσ = dΩ as a function of ϴ, α
and β (which is implicitly a function of ϴ from (a) and Snell's
Law).
c) The smallest value of ϴ is called the rainbow scattering angle.
Why? Find it numerically to first order in  if the index of
refraction is n = 1:333 +  .
d) The visual spectrum runs from violet, where n = 1:343, to red,
where n = 1:331. Find the angular radius of the rainbow's circle,
and the angular width of the rainbow, and explain whether the
red or blue is on the outside.
2. Consider the case of a projectile particle of mass μ being deflected
by a repulsive central force potential U(r) > 0. As the projectile
particle approaches from the right (at r =  , ϴ = 0) moving with
speed u, it is progressively deflected until it reaches a minimum
radius  at    after which the projectile particle moves away
from the repulsion center until it reaches r   at a deflection angle
ϴ and again moving with speed u. From the Figure shown below,
we can see that the scattering process is symmetric about the line of
closest approach.

105
PHY 301 CLASSICAL MECHANICS II

Define the angle  in terms of impact parameter b, μ and E (total


energy).

7.0 Further Reading and Other Resources

Hand, L. N. and Finch, J. D. (1998). Analytical Mechanics. Cambridge


University Press.

Thornton, S.T. and Marion, J. B. (2014.). Classical Dynamics of


Particles and Systems, Fifth Edition. Thomson Books/Cole

Goldstein, H., Poole, C., and Safko, J. (2000). Classical Mechanics,


Third Edition. Horoloma Book Jovanovich College Publisher.

Video Link:
https://www.youtube.com/playlist?list=PLtUBquogGkRukyAAWOmq2
LmQx8Yf9H_3E

106
PHY 301 CLASSICAL MECHANICS II

MODULE 4 MOTION IN NON-INERTIAL REFERENCE


FRAME

Unit 1 Time Derivative in Fixed and Rotating Frames


Unit 2 Motion Relative to Earth

UNIT 1 TIME DERIVATIVES IN FIXED AND ROTATING


FRAMES

CONTENTS

1.0 Introduction
2.0 Objectives
3.0 Main Contents
3.1 Time Derivatives in Fixed and Rotating Frames
3.2 Accelerations in Rotating Frames
3.3 Lagrangian Formulation of Non-Inertia Motion
4.0 Conclusion
5.0 Summary
6.0 Tutor Marked Assignment
7.0 Further Reading and Other Resources

1.0 Introduction

A non-inertial reference frame is a reference frame that is not tied to


the state of motion of an observer, with the property that each physical
law portrays itself in the same form in every inertial frame. As such, the
laws of physics in such a frame do not take on their most simple form,
as required by the special principle of relativity. To explain the motion
of bodies entirely within the viewpoint of non-inertial reference frames,
fictitious forces (also called inertial forces, pseudo-forces and
d'Alembert forces) must be introduced to account for the observed
motion, such as the Coriolis force or the centrifugal force, as derived
from the acceleration of the non-inertial frame

2.0 Objectives

By the end of this unit, you will be able to:

 Derive the time derivatives of vector A in fixed and rotating


reference frame.
 State the expressions for translational velocity and acceleration, V
and A respectively.
 Determine the relationship between the velocities of fixed and
rotating reference frame.

107
PHY 301 CLASSICAL MECHANICS II

 Show the relationship between the acceleration of fixed and


rotating reference frame.

3.0 Main Contents

3.1 Time Derivatives in Fixed and Rotating Frames

Let us consider the time derivative of an arbitrary vector A in two


reference frames. The first reference frame is called the fixed frame and
is expressed in terms of the Cartesian coordinates r  ( x; y; z) . The
second reference frame is called the rotating frame and is expressed in
terms of the Cartesian coordinates r = (x; y; z). In the Figure below, the
rotating frame shares the same origin as the fixed frame and the rotation
angular velocity ω of the rotating frame (with respect to the fixed frame)
has components (ωx, ωy, ωz).

figure 1.1
Since observations are often made in a rotating frame of reference, we
decompose the vector A in terms of components Ai in the rotating frame
(with unit vectors x̂ i ). Thus, A  A i x̂ i (using the summation rule) and
the time derivative of A as observed in the fixed frame is
dA dA i i dxˆ i
 xˆ  A i .
dt dt dt
(1.1)
The interpretation of the first term is that of the time derivative of A as
observed in the rotating frame (where the unit vector x̂i are constant)
while the second term involves the time dependence of the relation
between the fixed and rotating frames. We now express dxˆ i / dt as a
vector in rotating frame as
dxˆ i
 R ij xˆ j   ijk  k xˆ j ,
dt
(1.2)

108
PHY 301 CLASSICAL MECHANICS II

where R represents the rotation matrix associated with the rotating frame
of reference; this rotation matrix associated is anti-symmetric (Rij =-Rji)
and can be written in terms of the anti-symmetric tensor εijk (defined in
terms of the vector product A  B  A i B j  ijk x̂ k for two arbitrary vectors
A and B) as Rij   ijkk (see appendix) where ωk denotes the components
of the angular velocity ω in the rotating frame. Hence, the second term
in equation 1.1 above becomes
dxˆ i
A i  A i  ijk  k xˆ j    A .
dt
(1.3)
This time derivative of an arbitrary rotating frame vector A in a fixed
frame is, therefore expressed as
 dA   dA 
    A ,
 dt  f  dt  r
(1.4)
 
where d dt
f
denotes the time derivative as observed in the fixed (f)

 
frame while d dt denotes the time derivative as observed in the
r
rotating (R) frame. An application of this formular relates to the time
derivative of the rotation angular velocity ω itself. One can easily see
that
 d   d 
       .
 dt  f  dt r
(1.5)
Since the second term of equation 1.4 above vanishes for A = ω; the
time derivative of ω is, therefore, the same in both frames of reference
and is denoted  in what follows.

Self Assessment Exercise A


1. Show that the time derivative of ω is the same in both fixed and
rotating frame of reference

3.2 Accelerations in Rotating Frames

We now consider the general case of a rotating frame and fixed frame
being related by translation and rotation.

109
PHY 301 CLASSICAL MECHANICS II

figure 1.2
In the figure 1.2 above, the position of a point P according to the fixed
frame of reference is labeled r  , while the position of the same point
according to the rotating frame of reference is labeled r, and r  R  r,
where R denotes the position of the origin of the rotating frame
according to the fixed frame. Since the velocity of the point P involves
the rate of change of position, we must now be careful in defining which
   
time-derivatives operator, d dt or d dt that is to be used.
f r

The velocities of point P as observed in the fixed and rotating frames are
defined as
 dr    dr 
vf    and vr   
 dt  f  dt r
(1.6)
respectively. Using equation 1.4 the relationship between the fixed
frame and rotating frame velocity is expressed as
 dR   dr 
vf        V  vr    r ,
 dt  f  dt  f
(1.7)
 dR 
where V    denotes the translation velocity of the rotating-frame
 dt  f
origin (as observed in the fixed frame).
Using equation 1.7 above, we are now in a position to evaluate
expressions for the acceleration of point P as observed in the fixed and
rotating frames of reference
 dv f   dv 
a f    and ar   r 
 dt f  dt r
(1.8)
110
PHY 301 CLASSICAL MECHANICS II

respectively. Hence, using equation 1.7 we find


 dV   dvr   d   dr 
af        r   
 dt  f  dt  f  dt  f  dt  f
 A  (ar    vr )    r    (vr    r ),
(1.9)
or a f  A  ar  2  vr    r    (  r ),
(1.10)
 dV 
where A    denotes the translational acceleration of the rotating
 dt  f
frame origin (as observed in the fixed frame of reference). We can now
write an expression for the acceleration of point P as observed in the
rotating frame as
ar  a f  A    (  r )  2  vr    r ,
(1.11)
which represent the sum of the net inertial acceleration (af –A), the
centrifugal acceleration      r  (see figure below)

figure 1.3
the coriolis acceleration  2  vr

figure 1.4

and an angular acceleration term    r which depends explicitly on


the time dependence of the rotation angular velocity ω.
The centrifugal acceleration (which is directed outwardly from the
rotation axis) represents a familiar non-inertial effect in Physics. A less
familiar non-inertial effect is the Coriolis acceleration. The Figure 1.4

111
PHY 301 CLASSICAL MECHANICS II

above shows that an object falling inwardly also experiences an


eastward acceleration.

Self Assessment Exercise B


1. Given the figure 2.1 above determine the relationship between
acceleration between fixed and rotating reference frame.
2. Make sure you have a good understanding of differentiation and
integration

3.2 Lagrangian Formulation Of Non-Inertia Motion

The Lagrangian for a particle of mass m moving in a non-inertial


rotating frame (with its origin coinciding with the fixed-frame origin) in
the presence of the potential U(r) is expressed as
m
L(r , r)  r    r  U (r ),
2

2
(1.12)
where ω is the angular velocity vector and we use the formula
r    r  r  2  (r  r)   2 r 2  (  r ) 2  .
2 2

(1.13)
Using the Lagrangian (equation 2.1), we now derive the general Euler-
Lagrange equation for r. First, we derive an expression for the canonical
momentum
L
P  m(r    r ),
dr
(1.14)
d  L 
and    m(r    r    r) .
dt  r 
(1.15)
Next we derive the partial derivative
L
 U (r )  m  r    (  r ),
r
(1.16)
so that the Euler-lagrange equation are
mr  U (r )  m  r  2  r    (  r )
. (1.17)
Here, the potential energy term generates the fixed-frame acceleration,
 U  ma f and the Euler-Lagrange equation 2.7 yields equation 1.11

Self Assessment Exercise C


1. State the Lagrangian for a particle moving in a non-inertia rotating
frame and the corresponding Lagrange’s equation of motion

112
PHY 301 CLASSICAL MECHANICS II

4.0 Conclusion

This time derivative of an arbitrary rotating frame vector A in a fixed


frame is, therefore expressed as
 dA   dA 
        A , where d dt
 dt  f  dt  r
  f
denotes the time derivative

 
as observed in the fixed (f) frame, while d dt denotes the time
r
derivative as observed in the rotating (R) frame. An application of this
formular relates to the time derivative of the rotation angular velocity ω
itself. One can easily see that
 d   d 
       . Since the second
 dt  f  dt r
term of equation above vanishes for A = ω; the time derivative of ω is,
therefore, the same in both frames of reference and is denoted as  .

5.0 Summary

 A non-inertial reference frame is a reference frame that is not


tied to the state of motion of an observer, with the property that
each physical law portrays itself in the same form in every
inertial frame.
 The first reference frame is called the fixed frame and is
expressed in terms of the Cartesian coordinates r  ( x; y; z) .
 The second reference frame is called the rotating frame and is
expressed in terms of the Cartesian coordinates r = (x; y; z).
 This time derivative of an arbitrary rotating frame vector A in a
fixed frame is, therefore expressed as
 dA   dA 
      A
 dt  f  dt  r
 
where d dt
f
denotes the time derivative as observed in the
 
fixed (f) frame while d dt denotes the time derivative as
r
observed in the rotating (R) frame.
 The velocities of point P as observed in the fixed and rotating
frames are defined as
 dr    dr 
vf    and vr    .
 dt  f  dt r
 Expressions for the acceleration of point P as observed in the
fixed and rotating frames of reference is

113
PHY 301 CLASSICAL MECHANICS II

a f  A  ar  2  vr    r    (  r ), where
 dV 
A  denotes the translational acceleration of the rotating
 dt  f
frame origin.
 The Lagrangian for a particle of mass m moving in a non-inertial
rotating frame (with its origin coinciding with the fixed-frame
origin) in the presence of the potential U(r) is expressed as
m
L(r , r)  r    r  U (r ).
2

2
 The Euler-Lagrange’s equation are
mr  U (r )  m  r  2  r    (  r )
.

6.0 Tutor Marked Assignments (TMAs)

1. Given the following parameter of a rotating reference frame,


r=10t+3, ω=6t2+4t+5. Determine the following: i) Centrifugal
acceleration, ii) Coriolis acceleration, iii) Centripetal
acceleration.
2. For a fixed and rotating reference frame associated with translation
and rotation as depicted in fig 1.2 if R=5t3+2 Determine the velocity
and acceleration of the fixed reference frame?
3. Find the Lagrangian and the corresponding Lagrange’s equation of
motion for a particle of mass m moving in non-inertia reference
frame.

7.0 Further Reading and Other Resources

Hand, L. N. and Finch, J. D. (1998). Analytical Mechanics. Cambridge


University Press.

Thornton, S.T. and Marion, J. B. (2014.). Classical Dynamics of


Particles and Systems, Fifth Edition. Thomson Books/Cole

Goldstein, H., Poole, C., and Safko, J. (2000). Classical Mechanics,


Third Edition. Horoloma Book Jovanovich College Publisher.

Video Link:
https://www.youtube.com/playlist?list=PLtUBquogGkRukyAAWOmq2
LmQx8Yf9H_3E

114
PHY 301 CLASSICAL MECHANICS II

UNIT 2 MOTION RELATIVE TO EARTH

CONTENTS

1.0 Introduction
2.0 Objectives
3.0 Main Contents
3.1 Motion relative to earth
3.2 Free-Fall Problem
4.0 Conclusion
5.0 Summary
6.0 Tutor Marked Assignment
7.0 Further Reading and Other Resources

1.0 Introduction

We can now apply the expressions developed in unit 1 of this module


above to the important case of the fixed frame of reference having its
origin at the center of Earth (point O in the figure 2.1 below) and the
rotating frame of reference having its origin at latitude λ and the
longitude ψ (point O in the figure below). We notes that the rotation of
the Earth is now represented as ψ = ω and that   0 .

2.0 Objectives

By the end of this unit you will be able to:


 Represent the rotation of the earth in terms of fixed frame of latitude
angle λ and the azimuthal angle ψ.
 State the expression for acceleration of a point as observed in the
rotating frame O.
 State the expression for pure gravitational acceleration in the
rotating frame of the Earth.
 Solve related problems.

115
PHY 301 CLASSICAL MECHANICS II

3.0 Main Contents

3.1 Motion relative to earth

figure 2.1
We arrange the (x, y, z) axis of the rotating frame so that the z-axis is a
continuation

of the position vector R of the rotating-frame origin, i.e.
R  Rzˆ in the rotating frame (where R=6378km is the Earth radius
assuming a spherical Earth). When expressed in terms of the fixed-frame
latitude angle λ and the azimuthal angle ψ, the unit vector ẑ is
zˆ  cos (cosxˆ  sinyˆ)  sin zˆ .
(2.1)
Likewise, we choose the axis to be tangent to a great circle passing
through the North and South poles, So that
xˆ  sin  (cosxˆ  sinyˆ)  coszˆ .
(2.2)
Lastly, the y axis is chosen such that
yˆ  zˆ  xˆ   sinxˆ  cosyˆ  .
(2.3)
We now consider the acceleration of a point as observed in the rotating
frame O by writing equation 1.11 as
d 2r     (  r )  2  dr .
2
 g0 f  R f
dt dt
(2.4)
The first term g 0 f represent the pure gravitational acceleration due to
the gravitational pull of the Earth on point P (as observed in the fixed
located at Earth’s center) and is given as
GM
g0 f   r
r
3

(2.5)

116
PHY 301 CLASSICAL MECHANICS II

where r   R  r is the position of point P in the fixed frame and r is the


location of P in the rotating frame. When expressed in terms of rotating-
frame spherical coordinates ( , ,  ) :
r   sin  cosxˆ  sin yˆ   coszˆ ,
(2.6)
the vector r  is written as
r   R   cos zˆ   sin  (cosxˆ  sin yˆ ),
(2.7)
r   R 2  2R cos   2  .
3 32
and thus,
(2.8)
The pure gravitational acceleration is, therefore, expressed in the
rotating frame of the Earth as
 1   cos zˆ   sin  (cosxˆ  sin yˆ ) 
g 0 f  -g0 r  
 
1  2 cos   2 
32

(2.9)
where gor acceleration due to gravitational pull on the earth as observed
in the rotating frame of the earth g0r  GM 2  9.789m / s 2 and ε = r/R
R
which is much lesser than 1.
The angular velocity in the fixed frame is   ẑ where
2rad
  7.27  10 5 rad / s
24  3600sec
(2.10)
is the rotation speed of Earth about its axis. In the rotating frame, we
find
  (sin zˆ  cosxˆ) .
(2.11)
Because the position vector R rotates with the origin of the rotating
frame, its time derivative yield
R f    R  (R cos  ) yˆ
(2.12)
R    R    (  R)   2 R cos  (cos zˆ  sin xˆ )
f f

(2.13)
and thus the centrifugal acceleration due to R is
R      R   g cos  (cos zˆ  sin xˆ ) ,
0r

(2.14)
2 2
where ω R=0.0337m/s can be expressed in terms of pure gravitational
acceleration g 0 r as  2 R  g0 r , where  =3.4 x 10-3. We now define the
physical gravitational acceleration as
g  g 0 f      R  r 
 
 g0 r  1   cos2  zˆ   cos  sin  xˆ , 
(2.15)

117
PHY 301 CLASSICAL MECHANICS II

where terms of order ε have been neglected. A plumb line experiences a


small angular deviation  ( ) from the true vertical given as
g  sin 2
tan  ( )  rx  .
g rz (2   )   cos 2
(2.16)
The function exhibits a maximum at a latitude  defined as
cos 2   /(2   ), so that
 sin 2 
tan     1.7  10 3
(2   )   cos 2 2 1  
(2.17)
  
or   5.86arc min at     rad  45.050 , we now return to
4 4
equation 2.4, which is written to lowest order in ε and α as
d 2r dr
2
 gr zˆ  2 
dt dt
(2.19)
 ( x sin   z cos  ) yˆ  y (sin xˆ  cos zˆ) .
dr
where 
dt
(2.20)
Thus, we find three component of equation 2.19 written explicitly as
x  2 sin y 

y  2 (sin x  cos z)
z  gr  2 cos y 

(2.21)
A first integration of equation 2.21 yields
x  2 sin y  C x 

y  2 (sin x  cos z )  C y 

z  gr t  2 cos y  C z 
(2.22)
where (Cx, Cy, Cz) are constants defined from initial conditions (x0, y0,
z0) and ( x0 , y0 , z0 )
C x  x  2 sin y0 

C y  y 0  2 sin x0  cos z0 

C z  z0  2 cos y0 
(2.23)
A second integration of equation 2.22 yields
t
x(t )  x0  C x t  2 sin   ydt,
0
t t
y (t )  y0  C yt  2 sin   xdt  2 cos   zdt
0 0

(2.24)
118
PHY 301 CLASSICAL MECHANICS II

t
1 2
z (t )  z0  Cz t  gt  2 cos   ydt
2 0

Which can also be re-written as



x(t )  x0  C x t  x(t ) 

y (t )  y 0  C y t  y (t ) 

1
z (t )  z 0  C z t  gr t 2  z (t )
2 
(2.25)
Where the coriolis drift are
 1
t

x(t )  2 sin   y0t  C yt 2   ydt  ,
 2 0 
 1
t
  1 1 3
t

y(t )  2 sin   x0 t  C x t   xdt   2 cos   z 0 t  C z t  gr t   zdt 
 2
  2

 2 0   2 6 0 
, (2.26)
 1
t

z (t )  2 cos   y0t  C yt 2   ydt  .
 2 0 
Note that each Coriolis drift can be expressed as an infinite series in
power of ω and that all Coriolis effects vanish when ω = 0.

Self Assessment Exercise A


1. Derive the constant Cx, Cy and Cz and the coriolis drift.

3.2 Free-Fall Problem

As an example of the importance of Coriolis effects, we consider the


simple free fall problem,
where ( x0 , y0 , z0 )  (0,0, h) and ( x0 , y0 , z0 )  (0,0,0)
So that the constants (equation 2.23) are Cx = 0 = Cz and Cy = 2ωhcosλ
Substituting these constants into equation 2.25 and keeping only terms
up to the first order in ω, we find
x(t )  0 ,
(2.27)
1
y (t )  gr t 3 cos  ,
3
(2.28)
1
z (t )  h  gr t 2 .
2
(2.29)

119
PHY 301 CLASSICAL MECHANICS II

Hence, a free falling object starting from rest touches the ground z(T) =
0 after a time T  2h g after which the object has drifted eastward by
r

a distance of
1  cos  8h 3
y (T )  gr T 3 cos   .
3 3 gr
(2.30)
Self Assessment Exercise B
1. State one of the importance of the coriolis effects
2. Find the eastward drift of an object falling freely from a height of
100m with a latitude of 450

4.0 Conclusion

Expressing x,y,z of a rotating frame of reference having its origin at


latitude λ and the longitude ψ (point O in the figure 2.1 above). in terms
of the fixed-frame latitude angle λ and the azimuthal angle ψ, we derive
the x,y,z which are function of t as follows:

x(t )  x0  C z t  x(t ) 

y (t )  y 0  C y t  y (t )  where x(t ) , y(t ) and z (t ) are referred

1
z (t )  z 0  C z t  gr t 2  z (t )
2 
to as Coriolis drift and this can be applied to get the drift of an object
falling freely from a height.

5.0 Summary
 According to figure 2.1 arrange the (x, y, z) axis of the rotating
frame so that the z-axis is a continuation 
of the position vector R
of the rotating-frame origin, i.e. R  Rzˆ in the rotating frame
(where R=6378km is the Earth radius assuming a spherical Earth).
 When expressed in terms of the fixed-frame latitude angle λ and the
azimuthal angle ψ, the unit vector ẑ
is zˆ  cos (cosxˆ  sinyˆ)  sin zˆ while that of x and y-axes are
xˆ  sin  (cosxˆ  sinyˆ)  coszˆ , yˆ  zˆ  xˆ   sinxˆ  cosyˆ  .

 When expressed in terms of rotating-frame spherical coordinates


( , ,  ) :
r   sin  cosxˆ  sin yˆ   coszˆ .
 In the rotating frame, we find
  (sin zˆ  cosxˆ) .
 The angular velocity in the fixed frame is   ẑ where
2rad
  7.27  10 5 rad / s
24  3600sec
120
PHY 301 CLASSICAL MECHANICS II

 The pure gravitational acceleration is, therefore, expressed in the


rotating frame of the Earth as
 1   cos zˆ   sin  (cosxˆ  sin yˆ ) 
g 0 f  g 0 r  
  
1  2 cos   2
32

2 2
where g 0r = GM/R =9.789m/s and ε = r/R which is much lesser than 1.
 Because the position vector R rotates with the origin of the rotating
frame, its time derivative yield
R f    R  (R cos  ) yˆ .
 The centrifugal acceleration due to R is
      R   g cos  (cos zˆ  sin xˆ )
R 0r

 The physical gravitational acceleration as


g  g 0 f      R  r 
  
 g0 r  1   cos2  zˆ   cos  sin  xˆ .
 The coriolis drift are
 1
t

x(t )  2 sin   y0t  C yt   ydt  .
 2

 2 0 
 1
t
  1 2 1 3
t

y (t )  2 sin   x0t  Cxt   xdt   2 cos   z0t  Czt  gt   zdt 
 2
 
 2 0   2 6 0 

 1
t

z (t )  2 cos   y0t  C yt 2   ydt 
 2 0 
 As an example of the importance of Coriolis effects, we consider the
simple free fall problem, where ( x0 , y0 , z0 )  (0,0, h) .

6.0 Tutor Marked Assignments (TMAs)


7.0
8.0 Find the eastward drift of an object falling freely from a height of
100m with a latitude of 450

121
PHY 301 CLASSICAL MECHANICS II

7.0 Further Reading and Other Resources

Hand, L. N. and Finch, J. D. (1998). Analytical Mechanics. Cambridge


University Press.

Thornton, S.T. and Marion, J. B. (2014.). Classical Dynamics of


Particles and Systems, Fifth Edition. Thomson Books/Cole

Goldstein, H., Poole, C., and Safko, J. (2000). Classical Mechanics,


Third Edition. Horoloma Book Jovanovich College Publisher.

Video Link:
https://www.youtube.com/playlist?list=PLtUBquogGkRukyAAWOmq2
LmQx8Yf9H_3E

122
PHY 301 CLASSICAL MECHANICS II

Appendix

Category Typerface Example


Scalar Lower case, normal weight, italic s
 
Vector Lower case or upper case, normal a or A
weight, italic, with arrow

Vector coordinate Lower case, with arrow ai
Representation
component
Unit vector Normal weight, italic, with hat â
Tensor coordinate Upper case, A
representation
Tensor coordinate Upper case, roman, with Ai i 1 n
representation component subscripts
The following special symbols are defined:
 ij kronecker delta =1 when i=j.
=0 otherwise.
 ijk Levi-Civita symbol or tensor
=1 if i,j,k is an even
permutation of 1,2,3
=0 otherwise (two or more of
i,j,k are equal)
=-1 if i,j,k is an odd
permutation of 1,2,3
a. even permutation of 1,2,3 are: 123, 231, 312.
b. odd permutation of 1,2,3 are: 321, 213,132.
c. repeated indices: 112, 122 and so on
Coordinate system
Cylindrical
The three coordinate are
 y
  x2  y2   tan 1   zz
x
The inverse relations are
x   cos y   sin zz
The area elements are
dd ddz ddz
The volume element is
dddz
The line element is
ds 2  d 2   2 d 2  dz 2
The unit vectors are
ˆ  xˆ cos  yˆ sin  ˆ   xˆ sin   yˆ cos zˆ  zˆ

123
PHY 301 CLASSICAL MECHANICS II

Spherical
The three coordinates are:
 y z
r  x2  y2  y2   tan 1     cos1  
x r
The inverse relations are
x  r sin cos y  r sin sin  z  r cos
The area elements are
r sindrd rdrd r 2 sindd
The volume element is
r 2 sindrdd
The line element is
ds 2  dr 2  r 2 d 2  r 2 sin 2 d 2
The unit vectors are
rˆ  xˆ sin  cos  yˆ sin  sin   zˆ cos ˆ   xˆ sin   yˆ cos

ˆ  xˆ cos cos  yˆ cos sin   zˆ sin .

Vector and Tensor Definitions and Algebraic Identities


Dot and Cross Products 
  
a  b  ab cos(a, b )   ij ai b j  Ai B j
    
a  b  nˆAB(a, b )   ijk ei A j Bk  Ai B j ,
   
where (a , b ) is the angle from a to b and n̂ is the unit vector normal to
 
the plane defined by a , b , and the right-hand rule, and the secondary
expression using vector components hold in rectangular coordinates.

Algebraic Identities
   
          
a  b  c  b  c  a   c  a  b  ab c

  
 
    
a  b  c  a  c b  a  b c  
a  b  c  d   a  b  c  d   a  b  d c  b  c d   a  c b  d   a  d b  c 
          

a  b  c  d   a  b  d c  a  b  cd  ab d c  ab c d  acd b  b cd a
           

.

Note that, if any of the vectors are the gradient vector  , care must be
taken in how the above

expressions are written out to ensure  acts on the appropriate vectors.
Any two quantities that
commute in the above should be commuted as necessary to get

reasonable behavior of  . But in

124
PHY 301 CLASSICAL MECHANICS II

some cases, even that may not be sufficient and you will have to keep

track of which vector should be acted on by  . A good example is the
  
second line when b   . In the simple case of a being constant, one

simply needs to move the b in the first term:



     
 
  

a    c  a  c   a   c
But if a depends on position and does not give zero when acted on by

 , then the above must be read with care. One has to somehow
 
remember that  should not be allowed to act on a since it does not act

on a in the original expression. Since the above expression does not
correctly convey that meaning, it is better to abandon the vector notation.
The completely unambiguous way to write it, using index notation, is
 
a    c  i   a j  i c j   a j  j ci .
  
j j
 
The key point is that in the first term, a is in a dot product with c , but

 must be allowed to act
  
on c first, and not as   c .

125

You might also like