Vous êtes sur la page 1sur 56

Classical Mechanics [Taylor, J.R.

]
Solution Manual

Written by
JGSK

Last Updated
December 10, 2016
Contents

1 Newtons Laws of Motion 2

2 Projectiles and Charged Particles 3

3 Momentum and Angular Momentum 4

4 Energy 5

5 Oscillations 6

6 Calculus of Variations 7

7 Lagranges Equations 8

8 Two-Body Central-Force Problems 15

9 Mechanics in Non-inertial Frames 18

10 Rotational Motion of Rigid Bodies 19

11 Coupled Oscillators and Normal Modes 35

12 Nonlinear Mechanics and Chaos 39

13 Hamiltonian Mechanics 40

14 Collision Theory 53

15 Special Relativity 54

16 Continuum Mechanics 55

1
Chapter 1

Newtons Laws of Motion

1.1 No plans to do this chapter yet.

2
Chapter 2

Projectiles and Charged Particles

2.1 No plans to do this chapter yet.

3
Chapter 3

Momentum and Angular Momentum

3.1 No plans to do this chapter yet.

4
Chapter 4

Energy

4.1 No plans to do this chapter yet.

5
Chapter 5

Oscillations

5.1 No plans to do this chapter yet.

6
Chapter 6

Calculus of Variations

6.1 No plans to do this chapter yet.

7
Chapter 7

Lagranges Equations

7.1 Write down the Lagrangian for a projectile (subject to no air resistance) in terms of its Cartesian
coordinates (x, y, z), with z measured vertically upward. Find the three Lagrange equations and
show that they are exactly what you would expect for the equations of motion.

Solution. The Lagrangian can be easily identified and written as such:


L=T U
1 (7.1)
= m(x2 + y 2 + z 2 ) mgz
2
A direct application of the Euler-Lagrange equations yields:

L d L
= mx = 0 (7.2)
x dt x
L d L
= my = 0 (7.3)
y dt y
L d L
= mz = mg (7.4)
z dt z

All three of which are expected for a projectile in free fall. Note the negative sign on the z direction
is due to the chosen sign convention.

7.8 (a) Write down the Lagrangian L(x1 , x2 , x1 , x2 ) for two particles of equal masses, m1 = m2 = m,
confined to the x axis and connected by a spring with potential energy U = 21 kx2 . [Here x is the
extension of the spring, x = (x1 x2 l), where l is the springs unstretched length, and I assume
that mass 1 remains to the right of mass 2 at all times.] (b) Rewrite L in terms of the new variables
X = 21 (x1 +x2 ) (the CM position) and x (the extension), and write down the two Lagrange equations
for X and x. (c) Solve for X(t) and x(t) and describe the motion.

8
Classical Mechanics [Taylor, J.R.]
Solution Manual

Solution. (a) The Lagrangian is simply:

1 1
L= m(x1 2 + x2 2 ) k(x1 x2 l)2 (7.5)
2 2

(b) We note that we can manipulate the new variables X = 21 (x1 + x2 ) and x = (x1 x2 l), as

such:

1
X = (x1 + x2 ) and x = x1 x2 (7.6)
2
1
x1 =X + x
2 (7.7)
1
x2 =X x
2
We can then rewrite the Lagrangian from Equation (7.5) as:
 
1 1 1 1
L = m (X + x)2 + (X x)2 kx2
2 2 2 2
  (7.8)
1 1 1
= m 2X 2 + x2 kx2
2 2 2

The Lagrange equations can be written fairly easily, one for each coordinate.

L d L
(1) : = (7.9)
x dt x
1 2k
kx = mx x = x (7.10)
2 m
L d L
(2) : = (7.11)
X dt X
0 = 2mX X = 0 (7.12)

(c) Now we are required to solve for X(t) and x(t). We tackle the easier of the two first:

X = 0 X(t) = v0 t + X0 (7.13)

where we have taken the liberty to introduce two integration constants that depend upon the initial

conditions. As for the other coordinate x, we invoke the familiar SHM solution:
r !
2k 2k
x = x x(t) = A cos + (7.14)
m m

where we have introduced two other integration constants A and . The center of mass moves with

constant velocity since no external forces are acting on it. The extension of the spring undergoes

simple harmonic motion, which just means that the two masses are oscillating with respect to each

other.

9
Classical Mechanics [Taylor, J.R.]
Solution Manual

7.33 A bar of soap (mass m) is at rest on a frictionless rectangular plate that rests on a horizontal table.
At time t = 0, I start raising one edge of the plate so that the plate pivots about the opposite
edge with constant angular velocity , and the soap starts to slide toward the downhill edge. Show
that the equation of motion for the soap has the form x 2 x = g sin t, where x is the soaps
distance from the downhill edge. Solve this for x(t), given that x(0) = x0 . [You can easily solve the
homogeneous equation; for a particular solution try x = B sin t and solve for B.]

Solution. To obtain the equation of motion, we shall utilise the Lagrange equations. First, we write
the Lagrangian, noting that the motion of the soap consists of both translational and rotational
motion:

1 1 2
L= mx2 + I mgx sin t
2 2
1 1
= mx2 + mx2 2 mgx sin t (7.15)
2 2

Applying the Lagrange equations, we get

d L L
= mx = m 2 x mg sin t
dt x x
x 2 x = g sin t (shown) (7.16)

Note that Equation (7.16) is a second order, inhomogeneous differential equation. We first consider
the homogeneous equation, which has a simple harmonic motion solution.

xH (t) = A1 et + A2 et (7.17)

Now we consider the particular solution. We make a trigonometric ansatz xP = B sin t. Substitut-
ing into Equation (7.16),

2 B sin t 2 B sin t = g sin t (7.18)


g
B= (7.19)
2 2

We can then compose our general solution for x(t):

x(t) = xH + xP
g
= A1 et + A2 et + sin t (7.20)
2 2

10
Classical Mechanics [Taylor, J.R.]
Solution Manual

We have the initial conditions: x(0) = x0 and x(0) = 0, which yield us:

g
A1 + A2 = x 0 and (A1 A2 ) + =0 (7.21)
2

Using these conditions, we can solve for the integration constants A1 and A2 :

x0 g
A1 = (7.22)
2 4 2
x0 g
A2 = + (7.23)
2 4 2

Finally, we are able to construct our general equation as:

x g  t  x0 g  t g
0
x(t) = e + + e + sin t (7.24)
2 4 2 2 4 2 2 2
g g
= x0 cosh t sinh t + sin t (7.25)
2 2 2 2

7.40 The spherical pendulum is just a simple pendulum that is free to move in any sideways direction.

(By contrast a simple pendulum is confined to a single vertical plane.) The bob of a spherical

pendulum moves on a sphere, centered on the point of support with radius r = R, the length of the

pendulum. A convenient choice of coordinates is spherical polars, r, , , with the origin at the point

of support and the polar axis (zaxis) pointing straight down. The two variables and make a

good choice of generalized coordinates. (a) Find the Lagrangian and the two Lagrange equations.

(b) Explain what the equation tells us about the z component of angular momentum lz . (c) For

the special case that =const, describe what the equation tells us. (d) Use the equation to

replace by lz in the equation and discuss the existence of an angle 0 at which can remain

constant. Why is this motion called a conical pendulum? (e) Show that if = 0 + , with  small,

then oscillates about 0 in harmonic motion. Describe the motion of the pendulums bob.

Solution. (a) With reference to Figure 7.1, we can write the Lagrangian in spherical coordinates as:

1
L= m(R2 2 + R2 sin2 ) + mgR cos (7.26)
2

11
Classical Mechanics [Taylor, J.R.]
Solution Manual

Figure 7.1: Sketch of the problem

Then, applying Lagrange equation to the two generalized coordinates:

d L L
=
dt

mR2 = mR2 2 sin cos mgR sin (7.27)

d L L
=
dt

mR2 sin2 = lz = const (7.28)

We leave it to the reader to simplify Equation (7.27). As for the latter equation, we have noted that
L
the Lagrangian is independent of , such that = 0.

(b) Equation (7.28) tells us that the z component of the angular momentum is constant.

(c) If = const, then = 0, and Equation (7.27) reduces to:

g
= sin (7.29)
R

That is, we can describe the motion of the pendulum like a simple pendulum, at a fixed plane

characterized by = 0 , for some fixed 0 .

(d) We want to first express in terms of lz (using Equation (7.28), then substitute it into Equa-

12
Classical Mechanics [Taylor, J.R.]
Solution Manual

tion (7.27).
lz
= (7.30)
mR2 sin2
 2
lz g
= 2 2 sin cos sin
mR sin R
(7.31)
lz2 cos g
= 2 4 3 sin
m R sin R
To explore a possible stationary solution of , we consider = 0. Following from Equation (7.31),
we obtain:
lz2 cos g
= sin (7.32)
m R4 sin3
2 R
cos = k sin4 = k(1 cos2 )2 (7.33)

where we have introduced the constant k = gm2 R3 /lz2 . To analyze the solution, we substitute
x = cos , into Equation (7.33) where x [0, 1]. WolframAlpha gives us the following plot: We note

Figure 7.2: Plot of Equation (7.33)

that there is only one solution for which Equation (7.33) is satisfied. This would correspond to the
solution 0 , as specified in the question.

(e) Consider a perturbation, = 0 + . We can Taylor expand the following quantities about the
equilibrium position x0 , and also write down the derivatives of :

cos cos 0  sin 0 (7.34)

sin sin 0 +  cos 0 (7.35)

=  (7.36)

Before we substitute these equations into Equation (7.31), we recall that at equilibrium, Equa-
tion (7.32) is satisfied. That is, we have:
lz2 g sin4 0
= (7.37)
m R4
2 R cos 0

13
Classical Mechanics [Taylor, J.R.]
Solution Manual

which we can use to simplify our expression slightly later on. With the last 4 equations from above,
Equation (7.31) gives us:

g sin4 0 g
 (cos 0  sin 0 )(sin 0 +  cos 0 )3 (sin 0 +  cos 0 ) (7.38)
R cos 0 R
g sin4 0 g
= (cos 0  sin 0 )(sin3 0 3 sin4 0 cos 0 ) (sin 0 +  cos 0 ) (7.39)
R cos 0 R

After some simplification and ignoring terms in 2 , we obtain:

1 + 3 cos2 0
 
g
 =  (7.40)
R cos 0

This shows that the pendulum will oscillate in a simple harmonic motion, with angular frequency
g 1 + 3 cos2 0
 
approximately 2 = , about the equilibrium position 0 .
R cos 0

14
Chapter 8

Two-Body Central-Force Problems

1 2
8.13 Two particles whose reduced mass is interact via a potential energy U = 2 kr , where r is the

distance between them. (a) Make a sketch showing U (r), the centrifugal potential energy Ucf (r), and

the effective potential energy Ueff (r). (Treat the angular momentum l as a known, fixed constant.)

(b) Find the equilibrium separation r0 , the distance at which the two particles can circle each

other with constant r. [Hint: This requires that dUeff /dr be zero.] (c) By making a Taylor expansion

of Ueff (r) about the equilibrium point r0 and neglecting all terms in (r r0 )3 and higher, find the

frequency of small oscillations about the circular orbit if the particles are disturbed a little from the

separation r0 .

Solution. (a) A sketch of all three graphs are shown in the figure below.

Figure 8.1: Graphs of the potentials U (r), Ucf (r) and Ueff (r)

15
Classical Mechanics [Taylor, J.R.]
Solution Manual

We can write the potentials as such:

1 2
U (r) = kr (8.1)
2
l2
Ucf (r) = (8.2)
2r2
1 2 l2
Ueff (r) = U (r) + Ucf (r) = kr + (8.3)
2 2r2

dUeff
(b) The equilibrium separation by setting = 0. We obtain:
dr
dUeff l2
= kr0 3 = 0
dr r0
 2 1/4
l
r0 = (8.4)
k

(c) Let us define r = r0 + . The Taylor expansion can be performed about the equilibrium r0 ,

noting that the first derivative was calculated before, and is equal to 0 at r0 :

d2 Ueff (r)

2
Ueff (r) = Ueff (r0 ) +  + ...
dr2 r=r0 +
l2 3l2
   
1 2 2
kr0 + +  k +
2 2r02 (r0 + )4
l2 3l2 4
   
1 2 2 5
kr0 + +  k + (r +  r )
2 2r02 0 0

l2 3l2 3l2
   
1 2
kr0 + 2 + 2 k + 4 + 3 5 (8.5)
2 2r0 r0 r0

For small oscillations , we can ignore terms of 3 and above, and note that the effective potential

takes on the form of a harmonic oscillator potential (e.g. one of an elastic spring). That is, it takes
p
the form of U (x) = U (0) + 21 k 0 x2 , and the frequency of oscillation for such a system is = k 0 /m.

Making such an observation, we can deduce the spring constant k 0 and oscillation frequency :

3l2
 
k 0 = 2 k + 4 = 2k + 6k = 8k
r0
s s
k0 8k
= = (8.6)

where we have used the result from Equation (8.4) to rewrite r0 in terms of k, the central force

constant. Do not confuse this with the spring constant k 0 used in the analogy earlier.

16
Classical Mechanics [Taylor, J.R.]
Solution Manual

Figure 8.2: Transfer orbit for Problem 8.34

8.34 Suppose that we decide to send a spacecraft to Neptune, using the simple transfer described in

Example 8.6 (c.f. textbook pg 318). The craft starts in a circular orbit close to the Earth (radius 1
AU) and is to end up in a circular orbit near Neptune (radius about 30 AU). Use Keplers third law
to show that the transfer will take about 31 years.

Solution. The transfer orbit in question is depicted in Figure 8.2, with the radius of the Earth orbit
as rE and the radius of the Neptune orbit as rN .
a3 GM
Keplers third law is stated as:
3
= .
4 2
We know (from the figure) that the sum of the radii equals twice the semi-major axis a of the
elliptical orbit. Explicitly, we have

rE + rN = 2a a = 15.5 AU (8.7)

And so to get from P to P 0 :


r
1 1 4 2 a3
Time of transfer = =
2 2 GM
= 1.926 109 s ' 31 years (shown) (8.8)

Note that I used the following constants: 1 AU = 1.496 1011 m, G = 6.67 1011 Nm2 kg-2 ,
MSun = 1.989 1030 kg.

17
Chapter 9

Mechanics in Non-inertial Frames

9.1 No plans to do this chapter yet.

18
Chapter 10

Rotational Motion of Rigid Bodies

m r0 = 0, can be paraphrased to say that the position vector of the CM relative


P
10.1 The result that
to the CM is zero, and, in this form, is nearly obvious. Nevertheless, to be sure that you understand
the result, prove it by solving r = R + r0 for r0 and substituting into the sum concerned.

Solution. We start with the equation given in the question, and substituting into the sum as
suggested:

r = R + r0 r0 = r R (10.1)
X X X
m r0 = (m r ) R m (10.2)

where the sum is taken implicitly over . We also have the following expression for the position of
the center of mass:

1 X
R= (m r ) (10.3)
M
P
Using this relation, and noting that the term m is simply the total mass M , Equation (10.2)
becomes:
X
m r0 = RM RM = 0 (10.4)

We have thus shown that the position of the CM relative to itself is zero.

10.3 Five equal point masses are placed at the five corners of a square pyramid whose square base is
centered on the origin in the xy plane, with side L, and whose apex is on the z axis at a height H
above the origin. Find the CM of the five mass system.

19
Classical Mechanics [Taylor, J.R.]
Solution Manual

Solution. The (x, y, z) coordinates of the 4 corners are (L/2, L/2, 0), (L/2, L/2, 0), (L/2, L/2, 0),

(L/2, L/2, 0), while the coordinates of the apex is given by (0, 0, H). Let m be the mass of one

point mass. To calculate the CM of the system, we perform the following:


P  
m i xi
1 L L L L
xCM = = + =0 (10.5)
5m 5 2 2 2 2
P  
m i yi 1 L L L L
yCM = = + =0 (10.6)
5m 5 2 2 2 2
P
m i zi H
zCM = = (10.7)
5m 5

The coordinates of the CM is given by (0, 0, H/5).

10.4 The calculation of centers of mass or moments of inertia usually involves doing an integral, most

often a volume integral, and such integrals are often best done in spherical polar coordinates. Prove

that:
Z Z Z Z
2
dV f (r) = dr r d sin d f (r, , ).

[Think about the small volume dV enclosed between r and r + dr, and + d, and and + d.]

If the volume integral on the left runs over all space, what are the limits of the three integrals on

the right?

Figure 10.1: Taken from http://astro-learned.blogspot.sg/

20
Classical Mechanics [Taylor, J.R.]
Solution Manual

Solution. Referring to the figure, a small volume element dV can be approximated to a cuboid with

volume:

dV = (dr)(rd)(r sin d) = r2 sin drdd (10.8)

Substituting this into the integral, taking note of the independence (or dependence) of the integrand

with respect to the 3 variables, we obtain, trivially, the result:


Z Z Z Z
2
dV f (r) = dr r d sin d f (r, , ). (10.9)

Another approach to this question is by considering the Jacobian. We express the Cartesian

coordinates in terms of the spherical polar coordinates:

x = r sin cos , y = r sin sin , z = r cos . (10.10)

We then have the volume element as dV = dxdydz = J(r, , ) drdd, with J(r, , ) being:

x x x


r

y sin cos r sin sin r cos cos
y y
J(r, , ) = = sin sin r sin cos r cos sin
r
cos 0 r sin
z z z

r

= r2 sin (10.11)

This yields us the same result as above. The integration limits for a volume integral over all space

are: 0 r , 0 2, 0 .

10.5 A uniform solid hemisphere of radius R has its flat base in the xy plane, with its center at the

origin. Use the result of Problem 10.4 to find the center of mass.

Solution. Due to the inherent spherical symmetry of the object, one might expect the x and y

coordinates of the center of mass to be zero. Nevertheless, they can be computed by:
R R
x dm y dm
xCM = R =0 yCM = R =0 (10.12)
dm dm

21
Classical Mechanics [Taylor, J.R.]
Solution Manual

Let be the density of the hemisphere. The z coordinate can be computed in a similar fashion:
R R
z dm r cos dV
zCM = R = R
dm dV
RR R /2 R 2
0
r r2 dr 0 sin cos d 0 d
= RR R /2 R 2
0
r2 dr 0 sin d 0 d
(R4 /4) (1/2) (2) 3R
= 3
= (10.13)
(R /3) (1) (2) 8

10.6 (a) Find the CM of a uniform hemispherical shell of inner and outer radii a and b and mass M
positioned as in Problem 10.5. (b) What becomes of your answer when a = 0? (c) What if b a?

Solution. (a) Once again, by symmetry we have xCM = yCM = 0. We set up a similar integral to
compute zCM , but with modified limits:
R R
z dm r cos dV
zCM = R = R
dm dV
Rb R /2 R 2
r r2 dr 0 sin cos d 0 d
= a Rb R /2 R 2
a
r2 dr 0 sin d 0 d
1 4 4
4 (b a ) (1/2) (2)
= 1 3 3
3 (b a ) (1) (2)
 4 4

3 b a
= (10.14)
8 b3 a3
(b) Note that when a = 0, we are simply computing the CM of a uniform hemisphere (as in
3
Problem 10.4), with b = R. It is then no surprise that our answer here is 8 b, which corresponds
with our answer from the previous problem.
(c) When b a, the hemispherical shell becomes infinitely thin, with radius a. We would expect
that zCM approaches a/2. Now let us compute it. Note that here, a is simply a number, and so we
are only treating b as a variable as it tends towards a. Since we have the limit of [0]/[0], we can
use LHopitals rule:
3 b4 a4
 
zCM = lim
ba 8 b3 a3
 3
3 4b
= lim
8 ba 3b2
3 4a3
 
a
= 2
= (10.15)
8 3a 2

22
Classical Mechanics [Taylor, J.R.]
Solution Manual

10.9 The moment of inertia of a continuous mass distribution with density % is obtained by converting

m 2 into the volume integral 2 dm = 2 % dV . Find the moment of inertia


P R R
the sum Iz =

of a uniform circular cylinder of radius R and mass M for rotation about its axis. Explain why the

products of inertia are zero.

Solution. For continuous masses, we have:

Z Z h Z 2 Z R
Iz = % 2 dV = % 2 d d dz
0 0 0
 4
R
= % (h)(2) (10.16)
4

M
Since the density % = , we finally obtain the moment of inertia Iz :
R2 h

M R2
Iz = (10.17)
2

The products of inertia are zero because of cylindrical symmetry. More explicitly, we have, for

example:

Z
Ixy = % xy dV
Z h Z R Z 2
= % dz 2 d sin cos d
0 0 0
| {z }
=0
=0

10.15 (a) Write down the integral (as in Problem 10.9) for the moment of inertia of a uniform cube of
2 2
side a and mass M , rotating about an edge, and show that it is equal to 3Ma . (b) If I balance

the cube on an edge in unstable equilibrium on a rough table, it will eventually topple and rotate

until it hits the table. By considering the energy of the cube, find its angular velocity just before

it hits the table. (Assume the edge does not slide on the table.)

23
Classical Mechanics [Taylor, J.R.]
Solution Manual

Solution. (a) The details of this integral can be found in the textbooks Example 10.2. I shall

present the final answer here without proof:

Z a,a,a y 2 + z 2

xy xz
I= yx x2 + z 2 yz dV
0,0,0 zx zy x2 + y 2
2 5
41 a5 14 a5

3a
M
= 3 41 a5 2 5
3a 14 a5

a
41 a5 14 a5 2 5
3a

8 3 3
M a2
= 3 8 3 (10.18)
12
3 3 8

Hence we can directly read off the moment of inertia about an edge: I = 32 M a2 .

(b) When the cube is resting stably on the table, the height of the CM above the table is a/2.

When it is held at an unstable equilibrium (assuming balanced on an edge at a 45 angle), the


a a
height of the CM above the table is = . Hence we can calculate the change in the
2 tan 45 2
potential energy of the cube as it falls:
   
a a 1 1
P.E. = M g = M ga (10.19)
2 2 2 2

1 2
The corresponding gain in kinetic energy is K.E. = I , where is the angular velocity just
2
before it hits the table and I is the moment of inertia about an edge (calculated in (a)). Equating

the two energies together, and solving for :

3g
2 = ( 2 1) (10.20)
2a

10.23 Consider a rigid plane body or lamina, such as a flat piece of sheet metal, rotating about a point

O in the body. If we choose axes so that the lamina lies in the xy plane, which elements of the

inertia tensor I are automatically zero? Prove that Izz = Ixx + Iyy .

Solution. If the lamina is on the xy plane, then z = 0 automatically. Hence, products of inertia

containing z is thus automatically zero. Explicitly, Ixz = Izx = Izy = Iyz = 0. We will not be able

to deduce anything about Iyx and Ixy .

24
Classical Mechanics [Taylor, J.R.]
Solution Manual

Consider:
Z
Izz = x2 + y 2 dV
Z Z
= x2 + z 2 dV + y 2 + z 2 dV

= Iyy + Ixx (10.21)

We can perform the second step because z 2 = 0.

10.25 (a) Find all nine elements of the moment of inertia tensor with respect to the CM of a uniform
cuboid (a rectangular brick shape) whose sides are 2a, 2b and 2c in the x, y, z directions and whose
mass is M . Explain clearly why you could write down the off-diagonal elements without doing
any integration. (b) Combine the results of part (a) and Problem 10.24 to find the moment of
inertia tensor of the same cuboid with respect to the corner A at (a, b, c). (c) What is the angular
momentum about A if the cuboid is spinning with angular velocity around the edge through A
and parallel to the x axis?

Solution. (a)

Zc,Zb,Z a
2
y + z2

xy xz
I= yx x2 + z 2 yz dV (10.22)
c,b,a zx zy x2 + y 2

M
Identify the density to be = 8abc . We compute the diagonal elements first.
Z cZ bZ a
Ixx = y 2 + z 2 dxdydz
c b a
Z cZ b
M
= 2a(y 2 + z 2 ) dydz
8abc c b
Z c 3
M 2b
= + 2bz 2 dz
4bc c 3
M 2b2 c 2c3
 
M 2
= + = (b + c2 )
2c 3 3 3
M 2
Iyy = = (a + c2 )
3
M 2
Izz = = (a + b2 )
3

where the other two moments of inertia can be directly calculated, or inferred from Ixx (due to the
symmetry present).

25
Classical Mechanics [Taylor, J.R.]
Solution Manual

The off-diagonal elements are zero because: xy, yz, xz etc. are all odd functions in x, y, z.
Integrating from a to a of x, for example, will yield 0.

(b) We shift the origin by d = (a, b, c). Using the results from Problem 10.24, the modified
moment of inertia is given by:
2
b + c2
2
b + c2

0 0 ab ac
0 M 2 2
I = 0 a +c 0 +M ba a2 + c2 bc
3
0 0 a2 + b2 ca cb a + b2
2

2
4b + 4c2

3ab 3ac
M
= 3ba 4a2 + 4c2 3bc (10.23)
3
3ca 3cb 4a2 + 4b2


(c) If the angular velocity is parallel to the x axis, we can write = 0 . The angular
0
momentum is then given as the matrix multiplication:
2
4(b + c2 )

M
L = I = 3ba (10.24)
3
3ca

10.26 (a) Prove that in cylindrical polar coordinates, a volume integral takes the form
Z Z Z Z
dV f (r) = d d dzf (, , z).

(b) Show that the moment of inertia of a uniform solid cone pivoted at its tip and rotating about
its axis is given by the integral:
Z Z h Z 2 Z r
Izz = % dV 2 = % dz d d 2 ,
V 0 0 0

3 2
explaining clearly the limits of integration. Show that the integral evaluates to 10 M R . (c) Prove
3 2
also that Ixx = 20 M (R + 4h2 ).

Solution. (a) Using cylindrical polar coordinates, we have the following coordinate transformations:
x = cos , y = sin . Hence consider the Jacobian for a small volume element:
x x x


z

y cos sin 0
y y

J(, , z) = = sin cos 0 = (10.25)
z
0 0 1
z z z

z

26
Classical Mechanics [Taylor, J.R.]
Solution Manual

And so we have dV = d d dz, which then yields us the integral as required.

(b) If we were to use cylindrical coordinates, Izz will be calculated as:


Z Z
Izz = % x2 + y 2 dV = % 2 dV
V V

Utilising the result from part (a), we then have:


Z h Z 2 Z r
Izz = % dz d 3 d (10.26)
0 0 0

M
where we have the density: % = . The volume could have been obtained by considering
1/3 R2 h
RRR
the triple integral dV , but here we quote a direct result.

The limits of integration are as such:

h is the height of the cone.

2 is taken due to the cylindrical symmetry.


z
r = R is derived from considering a set of similar triangles in Figure 10.2, with R being
h
the radius of the base of the cone.

Let us now evaluate this integral.


h 2
R4 z 4
Z Z
Izz = % dz d
0 0 4h4
h
R4 z 4
Z
=% dz
0 2h4
3M R4 h
=
R2 h 10
3 2
= MR (shown) (10.27)
10

(c) Using the same notations as before, Ixx is calculated as:


Z Z
Ixx = % y 2 + z 2 dV = % 2 sin2 + z 2 dV
V V
Zh Z 2 Z r
=% dz d 3 sin2 + z 2 d
0 0 0
h 2 4 4
R2 z 4
Z Z
R z 2
=% dz sin + d
0 0 4h4 2h2
h
R4 z 4 R2 z 4
Z
=% 4
+ dz
0 4h h2
3
= M (R2 + 4h2 ) (shown)
20

27
Classical Mechanics [Taylor, J.R.]
Solution Manual

Figure 10.2: Similar triangles in a cone for Problems 10.26 and 10.27

10.27 Find the inertia tensor for a uniform, thin hollow cone, such as an ice-cream cone, of mass M ,

height h, and base radius R, spinning about its pointed end.

z
Solution. At any particular z, we have the corresponding radius r = R . We obtain this ratio by
h
considering a set of similar triangles, as seen in Figure 10.2. Using cylindrical coordinates (r, , z),

we first calculate the density % = M/A, where A is the curved surface area of the cone.
Z Z hZ 2
A= dA = r d dz
0 0

= Rh (10.28)
M
%= (10.29)
Rh

The inertia tensor is then given by:

y2 + z2

ZZ xy xz
I=% yx x2 + z 2 yz dA
zx zy x + y2
2
2 2
Z h Z 2 r sin + z 2 r2 sin cos rz cos
=% r2 sin cos r2 cos2 + z 2 rz sin r d dz (10.30)
0 0 rz cos rz sin r2

Note that off-diagonal elements equal to zero because


Z 2 Z 2 Z 2
sin d = cos d = sin cos d = 0
0 0 0

28
Classical Mechanics [Taylor, J.R.]
Solution Manual

Z 2 Z 2
Using the fact that sin2 d = cos2 d = , we have the diagonal components of the
0 0
inertia tensor to be:

h
R3 z 3 Rz 3
Z
Ixx = Iyy = % + 2 dz
0 h3 h
M 2
= (R + 2h2 ) (10.31a)
4

h
R3 z 3
Z
Izz = % 2 dz
0 h3
M R2
= (10.31b)
2

Finally we obtain the inertia tensor for a hollow cone to be:


2
R + 2h2

0 0
M
I= 0 R2 + 2h2 0 (10.32)
4
0 0 2R2

10.35 A rigid body consists of three masses fastened as follows: m at (a, 0, 0), 2m at (0, a, a) and 3m at

(0, a, a). (a) Find the inertia tensor I. (b) Find the principal moments and a set of orthogonal

principle axes.

Solution. (a) The inertia tensor of a rigid body with discrete masses is given by:
2
yi + zi2 xi yi

X xi zi
I= mi yi xi x2i + zi2 yi zi
i zi xi zi yi x2i + yi2

m(0) + 2m(2a2 ) + 3m(2a2 )



0 0
= 0 m(a2 ) + 2m(a2 ) + 3m(a2 ) 2m(a2 ) 3m(a2 )
2 2
0 2m(a ) 3m(a ) m(a2 ) + 2m(a2 ) + 3m(a2 )

10 0 0
= ma2 0 6 1
0 1 6

(b) We have to find directions of such that I = , where is the moment of inertia about

the principal axes. This eigenvalue problem has a non-trivial solution when det(I 1) = 0.

1
Firstly, by inspection, we know that 1 = 0 is an eigenvector, or equivalently, one of the
0
principal axes directions, with 1 = 10ma2 . To check that this is true, one only needs to substitute

29
Classical Mechanics [Taylor, J.R.]
Solution Manual

1 into I, and verify that one obtains 1 .


For the other two solutions, we just need to solve:

6ma2 ma2

ma2 =0
6ma2
(6ma2 2 )2 (ma2 )2 = 0

And so the eigenvalues are:

2 = 5ma2 and 3 = 7ma2

When 2 = 5ma2 , we have:


    
1 1 y 0
ma2 =
1 1 z 0
y = 1, z = 1

Normalizing, we obtain

0
1
2 = 1 (10.33)
2 1

Then when 3 = 7ma2 , we have, instead:


    
1 1 y 0
ma2 =
1 1 z 0
y = 1, z = 1

Normalizing, we obtain

0
1
3 = 1 (10.34)
2 1

Summarizing, with the set of orthogonal principle axes {1 , 2 , 3 }, the inertia tensor consisting
of the principal moments is given by:
10ma2

0 0
0
I = 0 5ma2 0 (10.35)
0 0 7ma2

10.36 A rigid body consists of three equal masses (m) fastened at the positions (a, 0, 0), (0, a, 2a), and
(0, 2a, a). (a) Find the inertia tensor I. (b) Find the principal moments and a set of orthogonal
principal axes.

30
Classical Mechanics [Taylor, J.R.]
Solution Manual

Solution. (a) We take reference from the previous problem (Problem 10.35). A similar, trivial

calculation can be performed to obtain the inertia tensor as:


2
yi + zi2 xi yi

X xi zi
I= mi yi xi x2i + zi2 yi zi
i zi xi zi yi x2i + yi2

10 0 0
= ma2 0 6 4
0 4 6

(b) Similar to the previous problem, we have to find directions of (principal axes), such that

we have I = , with being the moment of inertia about each principal axis. This eigenvalue

equation has non-trivial solutions if and only if det(I 1) = 0.

From theform of the inertia tensor above, we observe by inspection that the first eigenvector is
1
1 = 0, with an eigenvalue of 1 = 10ma2 .
0
For the other 2 axes, we consider the determinant:

6ma2 4ma2

4ma2 =0
6ma2

(6ma2 2 )2 (4ma2 )2 = 0

And so the eigenvalues are:

2 = 10ma2 and 3 = 2ma2

When 2 = 10ma2 , we have:


    
4 4 y 0
ma2 =
4 4 z 0
y = 1, z = 1

Normalizing, we obtain

0
1
2 = 1 (10.36)
2 1

Then when 3 = 2ma2 , we have, instead:


    
4 4 y 0
ma2 =
4 4 z 0
y = 1, z = 1

31
Classical Mechanics [Taylor, J.R.]
Solution Manual

(0, 1, 0)

dy

(0, 0, 0) (1, 0, 0)

Figure 10.3: Metal Triangle for Problem 10.37

Normalizing, we obtain

0
1
3 = 1 (10.37)
2 1

Summarizing, with the set of orthogonal principle axes {1 , 2 , 3 }, the inertia tensor consisting
of the principal moments is given by:
10ma2

0 0
0
I = 0 10ma2 0 (10.38)
0 0 2ma2

10.37 A thin, flat, uniform metal triangle lies in the xy plane with its corners at (1, 0, 0), (0, 1, 0) and the
origin. Its surface density (mass/area) is = 24. (a) Find the triangles inertia tensor I. (b) What
are its principal moments and the corresponding axes?

Solution. (a) The inertia tensor is given by:


2
y + z2

Z xy xz
I= yx x2 + z 2 yz dA
A zx zy x + y2
2
2
Z y xy 0
(z=0)
= yx x2 0 dA (10.39)
A 0 0 x + y2
2

We shall compute the elements of the inertia tensor component wise. Take reference from Fig-
ure 10.3, noting that the hypotenuse is characterized as: y = 1 x:
Z 1 Z 1y
Ixx = y 2 dx dy
0 0
Z 1
= y 2 (1 y) dy = 2 (10.40)
0
Z 1Z 1x
Iyy = x2 dy dx
0 0
Z 1
= x2 (1 x) dx = 2 (10.41)
0

32
Classical Mechanics [Taylor, J.R.]
Solution Manual

Z 1 Z 1y
Izz = x2 + y 2 dx dy
0 0

= Ixx + Iyy = 4 (c.f. Problem 10.23) (10.42)


Z 1 Z 1y
Ixy = Iyx = xy dx dy
0 0
1
(1 y)2
Z
= y dy = 1 (10.43)
0 2

Putting all the numbers together, we get the inertia tensor to be:

2 1 0
I = 1 2 0 (10.44)
0 0 4

(b) For the set of principal axes {1 , 2 , 3 }, we necessarily have Ii = i i for i = 1, 2, 3. Note

that i is the moment of inertia about the principal axis in the direction of i . The eigenvalue

equation has non-trivial solutions if


and only if det(I 1) = 0.
0
By inspection, we have that 3 = 0 is an eigenvector, with corresponding eigenvalue/ moment
1
of inertia 3 = 4. Verify this by ensuring the eigenvalue equation stated above is satisfied for this

choice of 3 .

As for the remaining two principal axes, we shall consider:



2 1

1 =0
2
(2 )2 1 = 0

And so the eigenvalues are:

1 = 1 and 2 = 3

When 1 = 1, we have:
    
1 1 x 0
=
1 1 y 0
x = 1, y = 1

Normalizing, we obtain

1
1
1 = 1 (10.45)
2 0

33
Classical Mechanics [Taylor, J.R.]
Solution Manual

Then when 2 = 3, we have:


    
1 1 x 0
=
1 1 y 0
x = 1, y = 1

Normalizing, we obtain

1
1
2 = 1 (10.46)
2 0

Summarizing, with the set of orthogonal principle axes {1 , 2 , 3 }, the inertia tensor consisting
of the principal moments is given by:

1 0 0
I 0 = 0 3 0 (10.47)
0 0 4

34
Chapter 11

Coupled Oscillators and Normal


Modes

11.5 (a) Find the normal frequencies, 1 and 2 , for the two carts shown in Figure 11.1, assuming that
m1 = m2 and k1 = k2 . (b) Find and describe the motion for each of the normal modes in turn.

Solution. (a) Let m1 = m2 = m, k1 = k2 = k, and rightwards be positive. We can write down the
forces acting on each cart:

F1 = k2 x2 + (k1 k2 )x1 = kx2 2kx1 (11.1)

F2 = k2 x2 + k2 x1 = kx2 + kx1 (11.2)

Using Newtons second law, we express the above two equations as a single matrix equation:
    
mx1 2k k x1
= (11.3)
mx2 k k x2

We look for the normal mode solutions that take the form x = Aeit , where x and A are column
vectors. Substituting into Equation (11.3),
    
2 A1 2k k A1
m = (11.4)
A2 k k A2

Figure 11.1: Two carts for Problems 11.5 and 11.6

35
Classical Mechanics [Taylor, J.R.]
Solution Manual

For non-trivial solutions, we have:



2k + m 2 k
=0
k k + m 2
k 2 3km 2 + (m 2 )2 = 0

We obtain the eigenfrequencies to be:



(3 5)
2
= 0 , (11.5)
2

where 0 = k/m. This works out to be 1 = 0.620 and 2 = 1.620 approximately.

(b) When = 1 , we have:


!
1 5

 
2 k k A1 0
1 5
=
k k A2 0
2

1+ 5
A1 = 1 and A2 = (11.6)
2

And so we have the amplitudes (the reader can normalize the expression if he so wishes) to be:
   
A1 1
= 1+5 (11.7)
A2 2

This means that both masses oscillate in phase, with the amplitude of mass 2 to be approximately
1.62 times that of mass 1.
When = 2 , we then have:
!
1+ 5

 
2 k k A1 0
1+ 5
=
k k A2 0
2

1 5
A1 = 1 and A2 = (11.8)
2

The amplitudes, when written in vector form is:


   
A1 1
= 1 5 (11.9)
A2 2

In this situation, both masses oscillate out of phase, with the amplitude of mass 2 being approxi-
mately 0.62 times that of mass 1.

11.6 Answer the same questions as in Problem 11.5 but for the case that m1 = m2 and k1 = 3k2 /2.
(Write k1 = 3k and k2 = 2k.) Explain the motion in the two normal modes.

36
Classical Mechanics [Taylor, J.R.]
Solution Manual

Solution. We use the same sign conventions as Problem 11.5, with m1 = m2 = m and k1 = 3k,
k2 = 2k. The forces on each mass is then:

F1 = k2 x2 + (k1 k2 )x1 = 2kx2 5kx1 (11.10)

F2 = k2 x2 + k2 x1 = 2kx2 + 2kx1 (11.11)

We then use Newtons second law to express the two equations above in a single matrix equation:
    
mx1 5k 2k x1
= (11.12)
mx2 2k 2k x2

Normal mode solutions: x = Aeit . Substituting into the above equation, we obtain:
    
2 A1 5k 2k A1
m = (11.13)
A2 2k 2k A2

For non-trivial solutions, we have the determinant vanishing:



5k + m 2 2k
=0
2k 2k + m 2
6k 2 7km 2 + (m 2 )2 = 0

We obtain the eigenfrequencies to be:

2 = 60 or 0 , (11.14)


where 0 = k/m. And so 1 = 60 and 2 = 0 .

When 1 = 60 , we have:
    
k 2k A1 0
=
2k 4k A2 0
A1 = 2 and A2 = 1 (11.15)

In vector form,
   
A1 2
= (11.16)
A2 1

With this normal mode, mass 1 and mass 2 are oscillating out of phase, with the amplitude of mass
1 being twice that of mass 2.
We do the same for 2 = 0 . We get:
    
4k 2k A1 0
=
2k k A2 0
A1 = 1 and A2 = 2 (11.17)

37
Classical Mechanics [Taylor, J.R.]
Solution Manual

In vector form,
   
A1 1
= (11.18)
A2 2

With this normal mode, mass 1 and mass 2 are oscillating in phase, with the amplitude of mass 1
being half that of mass 2.

38
Chapter 12

Nonlinear Mechanics and Chaos

12.1 No plans to do this chapter yet.

39
Chapter 13

Hamiltonian Mechanics

13.3 Consider the Atwood Machine of Fig 13.1, but suppose that the pulley is a uniform disc of mass M

and radius R. Using x as your generalized coordinate, write down the Lagrangian, the generalized

momentum p, and the Hamiltonian H = px L. Find Hamiltons equations and use them to find

the acceleration x.

x
y

m1

m2

Figure 13.1: The Atwood Machine

M R2
Solution. The moment of inertia of the disc is , and angular velocity of the disc is given by
2
x
= . Since the length of the string is fixed (say l), we can write the following equation to
R
incorporate the constraint:

y + x + R = l y = x + const (13.1)

40
Classical Mechanics [Taylor, J.R.]
Solution Manual

Now we can express the kinetic and potential energies in terms of a single coordinate x.
 2
1 2 1 M R2 x
T = (m1 + m2 )x +
2 2 2 R
 
1 1
= m1 + m2 + M x2 (13.2)
2 2
U = m2 gy m1 gx

= (m2 m1 )gx + const (13.3)

The Lagrangian is therefore:

L=T U
 
1 1
= m1 + m2 + M x2 + (m1 m2 )gx + const (13.4)
2 2

The generalized momentum p can be written as such:


 
L 1
p= = m1 + m2 + M x (13.5)
x 2

The Hamiltonian is then given by:

H = px L
p2 p2
 
1
= (m1 m2 )gx const
m1 + m2 + 12 M 2 m1 + m2 + 21 M
p2
 
1
= (m1 m2 )gx const (13.6)
2 m1 + m2 + 12 M

We note that H is just the total energy of the system, that is, H = T +U . The Hamiltons equations
are given by:

H
x =
p
p
= (13.7)
m1 + m2 + 12 M
H
p =
x
= (m2 m1 )g (13.8)

We can obtain the acceleration rather trivially from Hamiltons equations (Eqns 13.7 and 13.8).

p
x =
m1 + m2 + 12 M
(m2 m1 )g
= (13.9)
m1 + m2 + 12 M

41
Classical Mechanics [Taylor, J.R.]
Solution Manual

13.5 A bead of mass m is threaded on a frictionless wire that is bent into a helix with cylindrical polar
coordinates (, , z) satisfying z = c and = R, with c and R constants. The z axis points
vertically up and gravity vertically down. Using as your generalized coordinate, write down the
kinetic and potential energies, and hence the Hamiltonian H as a function of and its conjugate
momentum p. Write down Hamiltons equations and solve for and hence z. Explain your result
in terms of Newtonian mechanics and discuss the special case that R = 0.

Solution. In cylindrical coordinates, we have, in general, the kinetic energy T to be:

1  2 
T = m + 2 2 + z 2 (13.10)
2

Substituting in the constraints z = c and = R, we obtain


  1   2 
T = m R2 2 + c (13.11)
2

The potential energy is simply given by the gravitational potential energy:

U () = mgz = mgc (13.12)

We can write the Lagrangian L as:

1
m R2 + c2 2 mgc

L=T U = (13.13)
2

The next step is to find the canonical momentum p and substitute it into the general expression
for the Hamiltonian.

L
= m R2 + c2

p= (13.14)

H = p L (13.15)
p2 p2
 
= mgc (13.16)
mR2 (R2 + c2 ) 2m (R2 + c2 )
p2
= + mgc (13.17)
2m (R2 + c2 )

Now, we turn to Hamiltons equations by taking appropriate derivatives of the Hamiltonian.

H p
= = (13.18)
p m (R2 + c2 )
H
p = = mgc (13.19)

42
Classical Mechanics [Taylor, J.R.]
Solution Manual

We are able to solve for and z as such:

p gc
= = (13.20)
m (R2 + c2 ) R2 + c2
gc2
z = c = (13.21)
R2 + c2

Let us unwrap the helix into a 2-D plane. When go round the helix an angle of = 2, the
vertical height we have gained is z = 2c, while the horizontal distance travelled is 2R. That is,
we have the following:

g g sin

2c

2R
Figure 13.2: Unwrapped helix into a 2-D plane

where the angle tan is given by c/R. From this, we note that we can write Eqn 13.21 as:

z = g sin2 (13.22)

An object sliding down a slope of angle has a component of acceleration down the slope given
by atang = g sin . So the bead has a tangential acceleration atang along the wire. The vertical
component of this acceleration (along the z axis) is z = atang sin = g sin2 . This is precisely the
acceleration calculated in Equation (13.22).

If R = 0, then by Equation (13.21), z = g, as expected, since = /2.

13.6 In discussing the oscillation of a cart on the end of a spring, we almost always ignore the mass of the
spring. Set up the Hamiltonian H for a cart of mass m on a spring (force constant k) whose mass M
is not negligible, using the extension x of the spring as the generalized coordinate. Solve Hamiltons
p
equations and show that the mass oscillates with angular frequency = k/(m + M/3). That is,
the effect of the springs mass is to add M/3 to m. (Assume that the springs mass is distributed
uniformly and that it stretches uniformly.)

Solution. With reference to the figure, let L be the length of the spring, while be the mass per
unit length of spring.

43
Classical Mechanics [Taylor, J.R.]
Solution Manual

m
y(s)

s y(L) = x
ds
L

Consider an element A of spring of length ds and mass ds, at a distance s from the fixed end. Let

y(s) be the displacement of the element A, and y(L) = x be the displacement of mass m at s = L.

Assume that the spring stretches uniformly, i.e. y(s) = xs/L, with y(0) = 0, y(L) = x. We can

then express the velocity and kinetic energy of element A as:

xs
y =
L
1 x2 s2
TA = ( ds)y 2 = ds (13.23)
2 2L2

Thus, the kinetic energy of the spring is given by:

x2 L 2
Z
Tspring = s ds
2L2 0
1 1
= Lx2 = M x2 (13.24)
6 6

where in the last line we have identified the total mass of the spring as M = L.

With the kinetic energy of the cart as Tcart = 21 mx2 , we have the total KE:

1 1
T = Tspring + Tcart = M x2 + mx2
6 2
1
= meff x2 (13.25)
2

where meff = m + 13 M is the effective mass.

The potential energy can be identified to be U = 12 kx2 , and thus the Lagrangian is:

1 1
L= meff x2 kx2 (13.26)
2 2

We use the Lagrangian to derive the momentum in terms of the velocities, and then use the inverse

relation to obtain the Hamiltonian H(px , x), as such:

L
px = = meff x (13.27)
x
p2x 1
H = px x L = + kx2 (13.28)
2meff 2

44
Classical Mechanics [Taylor, J.R.]
Solution Manual

The Hamiltons equations of motions give us:

H px H
x = = px = = kx (13.29)
px meff x

Differentiating the first of the two equations with respect to time, and using the second equation,
we obtain:

px k
x = = x = 2 x (13.30)
meff meff

where 2 = k/meff . More explicitly, this is the equation of simple harmonic motion with frequency
r s
k k
= = (13.31)
meff m + M/3

13.17 Consider the mass confined to the surface of a cone described in Example 13.4 (page 533). We saw
that there are solutions for which the mass remains at the fixed height z = z0 , with fixed angular
velocity 0 , say. (a) For any chosen value of p , use (13.34) to get an equation that gives the
corresponding value of the height z0 . (b) Use the equations of motion to show that this motion
is stable. That is, show that if the orbit has z = z0 + , with  small, then  will oscillate about

zero. (c) Show that the angular frequency of these oscillations is = 30 sin , where is the
half angle of the cone (tan = c where c is the constant in = cz). (d) Find the angle for which
the frequency of oscillation is equal to the orbital angular velocity 0 , and describe the motion
in this case.

Solution. The two equations in (13.34) from the textbook (pg 534) are:

pz
z = (13.32)
m(c2 + 1)
p2
pz = mg (13.33)
mc2 z 3

The above equations were obtained from constructing the Lagrangian, and writing the Lagrange
equations. For completeness, the Hamiltonian of the system can be written as:
" #
1 p2z p2
H= + + mgz (13.34)
2m c2 + 1 c2 z 2

45
Classical Mechanics [Taylor, J.R.]
Solution Manual

Figure 13.3: A mass m constrained to move on the surface of a cone. Taken from textbook pg 533.

(a) To maintain at a fixed height, we require z = 0, which from Equation (13.32), we infer that it
is necessary that pz = 0. And so pz = 0. From Equation (13.33), we see that the height z0 which
corresponds to a particular value of p is:
!1/3
p2
z0 = (13.35)
m2 c2 g

(b) From Equation (13.32), we differentiate once with respect to time to get an equation of motion
for z as:

pz
z =
m(c2 + 1)
!
1 p2
= mg (13.36)
m(c2 + 1) mc2 z 3

We consider a perturbation of z about the equilibrium z0 , that is, z = z0 + , for small . We have
the following approximations:

z =  (13.37)
 

z 3 = z03 1 3 (13.38)
z0

Substituting these relations into Equation (13.36), we obtain:

3 p2
 =  (13.39)
m(c2 + 1) mc2 z04

p2
where we have used the fact that = mg (from Equation (13.35) in an intermediate step. To
mc2 z03
simplify Equation (13.39) further, we recall a relation between and p derived from the Hamiltons

46
Classical Mechanics [Taylor, J.R.]
Solution Manual

equations.
H p
= = (13.40)
p mc2 z 2
We use this to replace p in Equation (13.39):

32 c2
 =  (13.41)
c2 + 1

This takes on the form of a SHM equation. (shown)

(c) The angular frequency of the oscillation can be simply obtained from the SHM equation:
s
32 c2 c
= = 3 (13.42)
c2 + 1 2
c +1

Finally, we can note that the fraction above is none other than sin , where is defined in the

following way (c.f. Figure 13.3): tan = cz/z. Hence, we have shown that = 30 sin .


(d) In order for = 0 , we require sin = 1/ 3, or:

= 35.26 (13.43)

In this case, the masss rotational frequency equals its oscillation frequency. As a result, its orbit

is closed.

13.23 Consider the modified Atwood machine shown in Figure 13.4. The two weights on the left have

equal masses m and are connected by a massless spring of force constant k. The weight on the right

has mass M = 2m, and the pulley is massless and frictionless. The coordinate x is the extension

of the spring from its equilibrium length; that is, the length of the spring is le + x where le is

the equilibrium length (with all the weights in position and M held stationary). (a) Show that

the total potential energy (spring + gravitational) is just U = 12 kx2 (plus a constant that we can

take to be zero). (b) Find the two momenta conjugate to x and y. Solve for x and y, and write

down the Hamiltonian. Show that the coordinate y is ignorable. (c) Write down the four Hamilton

equations and solve them for the following initial conditions: You hold the mass M fixed with the

whole system in equilibrium and y = y0 . Still holding M fixed, you pull the lower mass m down a

distance x0 and at time t = 0 you let go of both masses. Describe the motion. In particular, find

the frequency with which x oscillates.

47
Classical Mechanics [Taylor, J.R.]
Solution Manual

Figure 13.4: Modified Atwood machine for Problem 13.23

Solution. (a) Let string length be L. We take downwards to be positive, and the origin to be the
centre of the pulley wheel. We also assume the radius of the pulley to be 0, without any loss in
generality. In addition, let us define the original, unstretched length of the spring (without load)
to be l0 . That is, le = l0 + mg/k, where k is the spring constant.

The two components of the potential energy can be written as:

Ugrav = mgy mg(y + le + x) 2m(L y)g

= mgle 2mLg mgx (13.44)


1
Uspring = k(x + mg/k)2 (13.45)
2

Summing the two potential energies give us:

(mg)2
 
1 2
U = (mgle 2mLg mgx) + kx + mgx +
2 2k
1 2
= kx + const (shown) (13.46)
2

Note that we can ignore the constant, and take it to be equal to zero. Henceforth, the potential
energy shall be computed only with U = 12 kx2 .

(b) We first complete the expression for the Lagrangian (L = T U ) by considering the kinetic
energy T of the system:

1 1 1
T = my 2 + (2m)y 2 + m(y + x)2 (13.47)
2 2 2

48
Classical Mechanics [Taylor, J.R.]
Solution Manual

Then, we can write the momentum conjugates as:


L T
px = = = m(y + x)
x x
(13.48)
L T
py = = = 3my + m(y + x) = m(4y + x)
y y
Solving the above equation for x and y, we find their expressions:
4px py
x =
3m
(13.49)
py px
y =
3m
The Hamiltonian can thus be written as:

H = xpx + ypy L
 
1 1 1
= (py px ) + px + kx2
2 2
(13.50)
2m 3 2
after some simplification. Note that this is simply the sum of the kinetic energies and potential
energy. Thus H represents the total energy of the system.
H
Since py = = 0, py is constant and thus y must be an ignorable coordinate. (shown)
y
(c) The four Hamilton equations are:
H H
px = = kx py = =0 (13.51)
x y
H 4px py H py px
x = = y = = (13.52)
px 3m py 3m
noting that the expressions in Equation (13.52) matches with those from Equation (13.49).

To solve these equations, we shall use the following initial conditions: x(0) = x0 , y(0) = y0 ,
x(0) = y(0) = 0. Furthermore, with Equation (13.48), we also have px (0) = py (0) = 0.
For the coordinate in x, we differentiate the expression for x to get:
4px py 4k
x = = x, (13.53)
3m 3m
where we have substituted the expressions for px and py from Equation (13.51). One can easily
recognise this taking the form of a simple harmonic motion, where the position and velocity is given
by:
r !
4k
x(t) = A cos t+ (13.54)
3m
r r !
4k 4k
x(t) = A sin t+ (13.55)
3m 3m

49
Classical Mechanics [Taylor, J.R.]
Solution Manual

where A and can be found by considering the initial conditions. It turns out that = 0 and
A = x0 , after a quick substitution of the initial conditions. As such, the mass attached to the
spring is undergoing simple harmonic motion described by:
r !
4k
x(t) = x0 cos t (13.56)
3m

r
4k
with angular frequency = t . As for the motion characterized by the coordinate y, since
3m
we know that (i) py is constant in time, and (ii) py (0) = 0, we have, necessarily, that py = 0 for all
time t. From the second equation in Equation (13.48), we have:
r r !
x 1 4k 4k
y = y = x0 sin t (13.57)
4 4 3m 3m

A quick integration (and applying initial conditions) yields us:


r !
1 1 4k
y(t) = (y0 + x0 ) x0 cos t (13.58)
4 4 3m

It is not surprising to see that y is executing simple harmonic motion as well.

13.25 Here is another example of a canonical transformation, which is still too simple to be of any real
use, but does nevertheless illustrate the power of these changes of coordinates. (a) Consider a
system with one degree of freedom and Hamiltonian H = H(q, p) and a new pair of coordinates Q

and P defined so that q = 2P sin Q and p = 2P cos Q. Prove that if qH = p and pH = q,
H H
it automatically follows that Q = P and P = Q. In other words, the Hamiltonian formalism
applies just as well to the new coordinates as to the old. (b) Show that the Hamiltonian of a
one-dimensional harmonic oscillator with mass m = 1 and force constant k = 1 is H = 12 (q 2 + p2 ).
(c) Show that if you rewrite this Hamiltonian in terms of the coordinates Q and P defined above,
then Q is ignorable. What is P ? (d) Solve the Hamiltonian equation for Q(t) and verify that, when
rewritten for q, your solution gives the expected behaviour.

Solution. (a) We have the following relation given in the question:



q = 2P sin Q
(13.59)
p = 2P cos Q

50
Classical Mechanics [Taylor, J.R.]
Solution Manual

We compute the relevant derivatives below:


q sin Q q
= = 2P cos Q (13.60)
P 2P Q
p cos Q p
= = 2P sin Q (13.61)
P 2P Q
H H
Consider the Hamiltonian equation for Q, recalling that = q and = p :
p q
H H p H q
= +
Q p Q q Q
p q
= q p
Q Q
   
q q p p p q
= Q + P Q + P
Q P Q Q P Q
   
q p p q q p p q
= Q + P
Q Q Q Q P Q P Q
 
q p p q
= 0 + P
P Q P Q
= P sin2 Q cos2 Q = P

(shown)

Similarly, we can consider the Hamiltonian equation for P :


H H p H q
= +
P p P q P
p q
= q p
 P P   
q q p p p q
= Q + P Q + P
Q P P Q P P
   
q p p q q p p q
= Q + P
Q P Q P P P P P
 
q p p q
= Q +0
Q P Q P
= Q cos2 Q + sin2 Q = Q

(shown)

(b) The Lagrangian for the 1-D harmonic oscillator is simply given by the quantity T U :
1 2 1 2
L=T U = mq kq (13.62)
2 2
We can express the velocity as a function of the momentum:
L p
p= = mq q = (13.63)
q m
Thus the Hamiltonian can be written as such:
1  p 2 1 2
H=T +U = m + kq
2 m 2
1 2
= (q + p2 ) (shown) (13.64)
2

51
Classical Mechanics [Taylor, J.R.]
Solution Manual

where the last line was obtained by setting k = m = 1, as given in the question.

(c) Using the coordinate transformations defined in Equation (13.59), we get:

1
2P sin2 Q + 2P cos2 Q = P

H= (13.65)
2

And so P is the Hamiltonian of the system. It can also be observed that Q is no longer present in
the Hamiltonian, and so, is an ignorable coordinate.

(d) We have from Hamiltons equations:

H
Q = =1 (13.66)
P

As such, we can easily find Q(t) by integrating, then solving for q:

Q=t+ (13.67)

q = 2 sin t + (13.68)

where is a constant of integration. This gives the expected behaviour (a simple harmonic oscil-
lation).

52
Chapter 14

Collision Theory

14.1 No plans to do this chapter yet.

53
Chapter 15

Special Relativity

15.1 No plans to do this chapter yet.

54
Chapter 16

Continuum Mechanics

16.1 No plans to do this chapter yet.

55

Vous aimerez peut-être aussi