Phy 301
Phy 301
GUIDE
PHY 301
CLASSICAL MECHANICS II
Lagos Office
14/16 Ahmadu Bello Way
Victoria Island, Lagos
e-mail: [email protected]
URL: www.nouedu.net
Printed 2023
ISBN: 978-058-005-0
Reviewed 2023
PHY 301 CLASSICAL MECHANICS II
Introduction
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.
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
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
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.
Appendix
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
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
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.
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
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.
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
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.
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
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
I wish you success in the course and I hope that you will find it both interesting
and useful.
MAIN
COURSE
CONTENTS
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
1
PHY 301 CLASSICAL MECHANICS II
2.0 Objectives
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
rk 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
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
3.2 Constraints
4
PHY 301 CLASSICAL MECHANICS II
5
PHY 301 CLASSICAL MECHANICS II
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
4.0 Conclusion
6
PHY 301 CLASSICAL MECHANICS II
5.0 Summary
7
PHY 301 CLASSICAL MECHANICS II
Video Link:
https://www.youtube.com/playlist?list=PLtUBquogGkRukyAAWOmq2
LmQx8Yf9H_3E
8
PHY 301 CLASSICAL MECHANICS II
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
2.0 Objectives
9
PHY 301 CLASSICAL MECHANICS II
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.
11
PHY 301 CLASSICAL MECHANICS II
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
12
PHY 301 CLASSICAL MECHANICS II
Dot Cancellation
ri ri
For holonomic constraints, it holds that .
qk qk
6.0 Tutor Marked Assignment
Video Link:
https://www.youtube.com/playlist?list=PLtUBquogGkRukyAAWOmq2
LmQx8Yf9H_3E
13
PHY 301 CLASSICAL MECHANICS II
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
2.0 Objectives
14
PHY 301 CLASSICAL MECHANICS II
(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)
xx zz 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
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 mgz .
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 21z1 and F22 r2 f 22z 21 do not vanish but the sum
does
2
F i2 ri f 21z1 f 22z 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 (m1z1 m2z 2 )
i 1 j 1 i 1
(3.14)
18
PHY 301 CLASSICAL MECHANICS II
ri m
r
dri dt i dq j
t j 1 q j
( nc ) ri
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 ) ri 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
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
r1 r2 z z
f z F11 F21 m1 g 1 m2 g 2
Z Z Z Z
m2 m1 g
(3.17)
4.0 Conclusion
(3.18)
(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
21
PHY 301 CLASSICAL MECHANICS II
Video Link:
https://www.youtube.com/playlist?list=PLtUBquogGkRukyAAWOmq2
LmQx8Yf9H_3E
22
PHY 301 CLASSICAL MECHANICS II
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
23
PHY 301 CLASSICAL MECHANICS II
But, we found earlier that we could write the virtual work as a sum over
only non-constraint forces, W Fij( nc ) ri .
(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 ) ri
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
(1.4)
T should be obtained by first writing T in terms of position velocities
ri 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 ri ri
mi ri pi ,
qk i qk i qk
(1.5)
T r r
mi ri i pi i ,
qk i qk 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 qk (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 qk 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
ql ql .
dt qk l ql qk ql qk tqk
(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 qk 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 ql qk 0 . Thus, we have
d ri ri ri ri
ql ,
dt qk qk
l q l t
q k
(1.9)
where the last step simply used the chain rule for evaluation of
ri 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 q1 would have made it
impossible to bring qk outside the sum because ql in general may
depend on ql (via Newton’s second law). So we now have
d T ri ri
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 ) ri 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
ma 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 ma2 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 z12 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)
28
PHY 301 CLASSICAL MECHANICS II
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.
4.0 Conclusion
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 qk qk
where T T qk , qk , 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
31
PHY 301 CLASSICAL MECHANICS II
Video Link:
https://www.youtube.com/playlist?list=PLtUBquogGkRukyAAWOmq2
LmQx8Yf9H_3E
32
PHY 301 CLASSICAL MECHANICS II
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
2.0 Objectives
33
PHY 301 CLASSICAL MECHANICS II
particle i.
34
PHY 301 CLASSICAL MECHANICS II
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 xi , yi , zi , 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
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 mx 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.
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.
37
PHY 301 CLASSICAL MECHANICS II
38
PHY 301 CLASSICAL MECHANICS II
39
PHY 301 CLASSICAL MECHANICS II
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
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)
41
PHY 301 CLASSICAL MECHANICS II
mr mr2 U (r ) .
(2.31)
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)
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
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)
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,
mr 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
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:
LQ, Q , t Lg (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.
46
PHY 301 CLASSICAL MECHANICS II
(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
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.
4.0 Conclusion
5.0 Summary
48
PHY 301 CLASSICAL MECHANICS II
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 2cos2 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
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.
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
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
x1 x 1l sin sin 1l cos cos
x x l sin sin l cos cos .
2 2 2
Because m11 m2 2 the cross terms cancel when we expand the
x-contribution to the kinetic energy.
m m2 2
1
2
1
m1 x12 m2 x22 1
2 2
x
m1m2l 2
2(m1 m2 )
sin sin cos cos
2
Similarly
m m2 2
1
2
1
m1 y12 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 z12 m2 z22 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
Video Link:
https://www.youtube.com/playlist?list=PLtUBquogGkRukyAAWOmq2
LmQx8Yf9H_3E
52
PHY 301 CLASSICAL MECHANICS II
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
53
PHY 301 CLASSICAL MECHANICS II
2.0 Objectives
54
PHY 301 CLASSICAL MECHANICS II
55
PHY 301 CLASSICAL MECHANICS II
56
PHY 301 CLASSICAL MECHANICS II
L L L
dH pi dqi qi dpi dqi dqi dt
i qi qi t
(3.5)
The derivative L qi 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 dqi qi dpi p i dqi pi dq ) dt
i t
L
qi 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
qi , 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 mxx 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
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
58
PHY 301 CLASSICAL MECHANICS II
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 m0 x 2
H .
2m 2
(3.15)
From equation 3.9 the Hamiltonian equations of motion are
p
x , p m0 x .
m
(3.16)
Gauge-transformed Harmonic Oscillator.
Now consider the gauge-transformed harmonic oscillator Lagrangian
mx 2 20 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 m0 x
2
1
m02 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).
60
PHY 301 CLASSICAL MECHANICS II
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
(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.
4.0 Conclusion
5.0 Summary
H H qk
, pk , t .
63
PHY 301 CLASSICAL MECHANICS II
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
p2 p2 p2
2 2 p U ( , )
R R cos2 2 R 2 2 R 2 cos2
p2 p2
p U ( , )
2 R 2 2 R 2 cos2
64
PHY 301 CLASSICAL MECHANICS II
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
Lq1 , q 2 , q1 , q 2 q12 k1 q12 k 2 q1 q 2 .
a bq12
Find the equations of motion using the Hamiltonian formulation.
Video Link:
https://www.youtube.com/playlist?list=PLtUBquogGkRukyAAWOmq2
LmQx8Yf9H_3E
65
PHY 301 CLASSICAL MECHANICS II
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
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)
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 ra mb rb ) . (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 ra mb rb .
(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)
68
PHY 301 CLASSICAL MECHANICS II
let’s re-write this in terms of rab
L rab ra rab rb rab rab 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
rab 2 rab 2 ab 2 rab 2 sin 2 abab 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.
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
(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.
70
PHY 301 CLASSICAL MECHANICS II
71
PHY 301 CLASSICAL MECHANICS II
72
PHY 301 CLASSICAL MECHANICS II
T
1
2
x1 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 ) .
2r 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
73
PHY 301 CLASSICAL MECHANICS II
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
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 2r
l
2
du 2 1
E u U .
2
2 d u
(1.37)
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
77
PHY 301 CLASSICAL MECHANICS II
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.
78
PHY 301 CLASSICAL MECHANICS II
5.0 Summary
1 2 l2 1
E r 2 U (r ) r 2 U eff (r ) ,
2 2r 2
l2
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
2 r d 2r 2
2 d u
80
PHY 301 CLASSICAL MECHANICS II
Video Link:
https://www.youtube.com/playlist?list=PLtUBquogGkRukyAAWOmq2
LmQx8Yf9H_3E
81
PHY 301 CLASSICAL MECHANICS II
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
82
PHY 301 CLASSICAL MECHANICS II
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.
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
,
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 2px 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
GM 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
GM 2 GM
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).
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
GM
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
GM
2 2
l l
E min U eff (r p ) ,
2 p 2 p 2 p 2
1 GM
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.
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.
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.
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
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.
93
PHY 301 CLASSICAL MECHANICS II
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
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.
Video Link:
https://www.youtube.com/playlist?list=PLtUBquogGkRukyAAWOmq2
LmQx8Yf9H_3E
95
PHY 301 CLASSICAL MECHANICS II
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
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
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
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 2bdb 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.
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.
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)
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
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
GM 2E
GM
We have v b 2 1 .
2E /
(3.16)
GM
that is b 2 1 .
2E
(3.17)
So we may now find b(ϴs):
GM GM
b csc2 s 1 cot s ,
2E 2 2E 2
(3.18)
s
where we take positive square root because 0 .
2 2
db
We will need to calculate the differential cross section, so, taking
d s
the derivative:
db 1 GM b
csc2 s csc s sec s .
d s 2 2E 2 2 2 2
(3.19)
102
PHY 301 CLASSICAL MECHANICS II
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.
4.0 Conclusion
5.0 Summary
103
PHY 301 CLASSICAL MECHANICS II
104
PHY 301 CLASSICAL MECHANICS II
105
PHY 301 CLASSICAL MECHANICS II
Video Link:
https://www.youtube.com/playlist?list=PLtUBquogGkRukyAAWOmq2
LmQx8Yf9H_3E
106
PHY 301 CLASSICAL MECHANICS II
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
2.0 Objectives
107
PHY 301 CLASSICAL MECHANICS II
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 ijkk (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.
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
figure 1.3
the coriolis acceleration 2 vr
figure 1.4
111
PHY 301 CLASSICAL MECHANICS II
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
mr 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
112
PHY 301 CLASSICAL MECHANICS II
4.0 Conclusion
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
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
mr U (r ) m r 2 r ( r )
.
Video Link:
https://www.youtube.com/playlist?list=PLtUBquogGkRukyAAWOmq2
LmQx8Yf9H_3E
114
PHY 301 CLASSICAL MECHANICS II
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
2.0 Objectives
115
PHY 301 CLASSICAL MECHANICS II
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 (cosxˆ sinyˆ) 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 (cosxˆ sinyˆ) coszˆ .
(2.2)
Lastly, the y axis is chosen such that
yˆ zˆ xˆ sinxˆ cosyˆ .
(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
(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
(2.24)
118
PHY 301 CLASSICAL MECHANICS II
t
1 2
z (t ) z0 Cz t gt 2 cos ydt
2 0
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.
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
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 (cosxˆ sinyˆ) sin zˆ while that of x and y-axes are
xˆ sin (cosxˆ sinyˆ) coszˆ , yˆ zˆ xˆ sinxˆ cosyˆ .
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) .
121
PHY 301 CLASSICAL MECHANICS II
Video Link:
https://www.youtube.com/playlist?list=PLtUBquogGkRukyAAWOmq2
LmQx8Yf9H_3E
122
PHY 301 CLASSICAL MECHANICS II
Appendix
123
PHY 301 CLASSICAL MECHANICS II
Spherical
The three coordinates are:
y z
r x2 y2 y2 tan 1 cos1
x r
The inverse relations are
x r sin cos y r sin sin z r cos
The area elements are
r sindrd rdrd r 2 sindd
The volume element is
r 2 sindrdd
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
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 cd ab d c ab c d acd b b cd 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