Vous êtes sur la page 1sur 62

Physics 225 Lecture Notes

A. A. Louro

Winter 2002
2
Contents

1 Rotations 1
1.1 Rotation kinematics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.1.1 Angular displacement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.1.2 Angular velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.1.3 Angular acceleration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.1.4 Are angular variables vectors? . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Relation between rotational and translational variables . . . . . . . . . . . . . . . . . 3
1.3 Rotational inertia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3.1 Rotational kinetic energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3.2 General procedure for calculating moments of inertia . . . . . . . . . . . . . . 6
1.3.3 Dependencies of the moment of inertia . . . . . . . . . . . . . . . . . . . . . . 6
1.3.4 The parallel-axis theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.4 Newton’s 2nd. law for rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.5 Rolling motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.6 Conservation of energy, including rotation . . . . . . . . . . . . . . . . . . . . . . . . 9
1.7 General definition of torque . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.8 Angular momentum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.8.1 Extension to systems of particles . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.9 Conservation of angular momentum . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2 Oscillations 15
2.1 Oscillators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.1.1 Simple harmonic motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.1.2 Hooke’s law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.1.3 Period and frequency of the oscillation . . . . . . . . . . . . . . . . . . . . . . 18
2.2 The ideal pendulum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.3 Elastic potential energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

i
ii CONTENTS

3 Waves 21
3.1 Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.1.1 Definitions and terminology . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.1.2 Why study waves? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.1.3 A general description of a sine wave . . . . . . . . . . . . . . . . . . . . . . . 23
3.1.4 Transverse and longitudinal waves . . . . . . . . . . . . . . . . . . . . . . . . 23
3.2 Waves in a string . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.2.1 The speed of propagation of a wave in a string . . . . . . . . . . . . . . . . . 23
3.2.2 Energy transport by a wave . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.3 The superposition principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.4 Interference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.5 Standing waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.6 Resonant modes of a string . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.7 Intensity of sound and sound level . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.7.1 The inverse square law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.7.2 Intensity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.7.3 The decibel scale . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.7.4 Hearing curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.8 The Doppler effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.8.1 Moving detector, stationary source . . . . . . . . . . . . . . . . . . . . . . . . 29
3.8.2 Moving source, stationary detector . . . . . . . . . . . . . . . . . . . . . . . . 30
3.8.3 The general Doppler effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

4 Fluids 33
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
4.1.1 Solids, liquids, and gases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
4.1.2 Density and pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
4.2 Fluid statics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
4.2.1 Variation of pressure with depth . . . . . . . . . . . . . . . . . . . . . . . . . 35
4.2.2 Buoyancy and Archimedes’ Principle . . . . . . . . . . . . . . . . . . . . . . . 36
4.3 Fluid dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
4.3.1 The velocity field in a fluid . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
4.3.2 Streamlines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.3.3 The continuity equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.3.4 Bernoulli’s law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

A Derivations for the final exam 43


A.1 Derivations for the final exam . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
CONTENTS iii

B The vector cross product 45


B.1 The vector cross product . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
B.1.1 Definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
B.1.2 The cross product is non-commutative . . . . . . . . . . . . . . . . . . . . . . 46
B.1.3 The x, y, and z components of a cross product . . . . . . . . . . . . . . . . . 46

C Dimensional analysis 49
C.1 The magic of dimensional analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
C.1.1 Step 1: What does the quantity of interest depend on? . . . . . . . . . . . . . 50
C.1.2 Step 2: Narrow down the possible functions by demanding that the dimen-
sions are right . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

D Average value of cos2 (x) 53


2
D.1 Average value of cos (x) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

E Work and potential energy 55


E.1 Work and potential energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
iv CONTENTS
Chapter 1

Rotations

1
2 CHAPTER 1. ROTATIONS

1.1 Rotation kinematics


We begin our study of rotation by considering a rigid body rotating about an axis fixed in space.
Figure 1.1 shows an example, captured at two instants in time. To describe the rotational motion,

s
Rotation axis
θ
r

t t+ ∆ t

Figure 1.1: Rotating rigid body.

we introduce new kinematic variables, analogous to displacement, velocity, and acceleration in


translational motion.

1.1.1 Angular displacement


During the interval [t, t + ∆t] the body has rotated through an angle θ. This is the angular
displacement. To make life simpler later, let’s agree from the start to measure all angles in
radians. Remember that one full turn is equivalent to an angle of 2π radians.

1.1.2 Angular velocity


The angular velocity is the rate at which the direction of the body changes as it spins:


ω= (1.1)
dt
(ω is the Greek letter “omega”). Angular velocity is measured in radians per second.

1.1.3 Angular acceleration


The angular acceleration is the rate at which the angular velocity changes over time:


α= (1.2)
dt
and it is measured in (rad/s)/s .
1.2. RELATION BETWEEN ROTATIONAL AND TRANSLATIONAL VARIABLES 3

1.1.4 Are angular variables vectors?


Are angular variables vectors? Sort of. For now, since we are only considering rotations about a
fixed axis, we don’t have to worry about this. However, θ, ω, and α are signed quantities. The
convention is that if we are looking down on the body, so that the rotation axis is perpendicular to
the page, and the angle/rotation sense/angular acceleration is counterclockwise, then the quantity
is negative.

1.2 Relation between rotational and translational variables


Consider any point on the rotating object at a distance r from the rotation axis. (For example,
the tip of the triangular object in Figure 1.1). As the body rotates through an angle θ, the point
describes a circular arc of length s. Since θ is measured in radians we can write

s = rθ (1.3)

The speed of the point is


ds
v= (1.4)
dt
and substituting s from equation (1.3), we find the following simple relation between the ordinary
speed v of the point, and the angular speed of the rotating body, ω:

d(rθ) dθ
v= =r = ωr (1.5)
dt dt
since r is of course constant. (The body is rigid, so the distance from the point to the rotation axis
doesn’t change).
If the body does not rotate at a constant rate, but rather has an angular acceleration α, the
point has an acceleration in the direction it is moving, i.e. tangent to its circular path. This
tangential acceleration is easily found:

dv d(rω)
at = = = rα (1.6)
dt dt
Recall that at the same time, because the point is turning in a circle, it has a centripetal, or
radial acceleration
v2
ar = (1.7)
r
which we can also write
ar = ω 2 r (1.8)
See Figure 1.2.
4 CHAPTER 1. ROTATIONS

a_t

a_r

Figure 1.2: Radial and tangential components of the acceleration

1.3 Rotational inertia


It’s becoming apparent that there is a close parallel between rotational and translational quantities.
There’s more; for example, just as mass measures the inertia of an object, that is how hard it is to
accelerate given a certain applied force, there is an equivalent concept in the case of rotation.
Since mass is defined in the context of Newton’s 2nd. law, we would look for the rotational
equivalent of Newton’s 2nd. law. There is indeed such a law, but we have to wait before we
introduce it. To discover what the rotational inertia is, we’ll try a different approach: We’ll use the
kinetic energy.

1.3.1 Rotational kinetic energy


The kinetic energy of a mass m moving with speed v is

1
K = mv 2 (1.9)
2

A rotating body also has kinetic energy, even though it isn’t going anywhere. Every atom in the
body is moving around the rotation axis as the entire object spins, so its total rotational kinetic
energy is simply the sum of the kinetic energies of all the particles that make up the object.
1.3. ROTATIONAL INERTIA 5

Following the analogy with translational motion, we write the rotational kinetic energy as
1
Krot = Iω 2 (1.10)
2
where I stands for the rotational equivalent of mass, as yet undetermined.

Exercise 1.3.1 Show that the units of I are kg m2 .

Consider for example a thin ring of mass M spinning about an axis through its centre, perpendicular
to the plane of the ring. (See Figure 1.3). If we isolate a tiny mass element m of the ring, its kinetic

Tiny mass element

Figure 1.3: Spinning ring.

energy is
1 1
mv 2 = mω 2 R2 (1.11)
2 2
Since all the mass elements that make up the ring are at the same distance R from the axis, and
rotate with the same ω, the total kinetic energy of the ring is
X 1 1 X 1
Krot = mω 2 R2 = ω 2 R2 m = M ω 2 R2 (1.12)
0
2 2 0
2
all m s all m s

Comparing this with equation (1.10), we see that

I = M R2 (1.13)

I is called the moment of inertia of the ring about this particular rotation axis. In general, if
the ring were rotating about any other axis, for example around a diameter, the moment of inertia
would be different because it depends on how the mass is distributed relative to the rotation axis.
6 CHAPTER 1. ROTATIONS

1.3.2 General procedure for calculating moments of inertia


Since the bodies we consider are continuous, the “tiny mass elements” that we isolate to calculate
the kinetic energy are really infinitesimals dm. The “sum over tiny mass elements” is in fact an
integral: Z Z
1 2 1 2
K= v dm = ω r2 dm (1.14)
2 2
where r is the distance of each mass element dm from the axis. This last integral is the moment of
inertia: Z
I = r2 dm (1.15)

In most cases, this integral turns out to be beyond the scope of this course. Refer to Table 11-2,
page 227 in the textbook for moments of inertia of some common bodies.

1.3.3 Dependencies of the moment of inertia


Unlike mass, which is a simple property of any object, independent of its shape, the moment of
inertia depends on several things:

• the mass of the object,

• the rotation axis

• the shape of the object, that is, how its mass is distributed around the rotation axis

1.3.4 The parallel-axis theorem


Very often we know the moment of inertia of an object when it rotates around an axis that goes
through the centre of mass (CM). The following theorem enables us to use this information to find
easily the moment of inertia about an axis parallel to the first one. Figure 1.4 shows an example.
The parallel axis is labelled “2”, and it is a distance h away from the centre of mass CM. We will
show that the moment of inertia with respect to axis 2 is related to the moment of inertia with
respect to the axis through CM by
I2 = ICM + M h2 (1.16)
where M is the mass of the object. Before we proceed with the proof, we need to be aware of
the definition of the centre of mass of an extended object. To find the coordinates of the centre
of mass, consider an isolated mass element dm with coordinates (x, y). A word equation for the
coordinates of the centre of mass goes something like this:
1.3. ROTATIONAL INERTIA 7

CM h 2

Figure 1.4: The parallel axis theorem.

[sum over all mass elements of (mass × position)] / (total mass) . This being a continuous
body, the sum is really an integral:
R
x dm
xCM = (1.17)
R M
y dm
yCM = (1.18)
M
Now we’re ready to tackle the parallel-axis theorem. As usual, to find a moment of inertia we isolate
first a mass element dm. For notation, see Figure 1.5. Notice in particular that the coordinate
system is chosen with CM at the origin, that is xCM = yCM = 0. We have
Z Z
ICM = r2 dm = (x2 + y 2 ) dm (1.19)

and Z Z Z
I2 = r22 dm = 2 2
[(h − x) + y ] dm = (x2 + y 2 + h2 − 2xh) dm (1.20)
so Z Z Z
2 2 2
I2 = (x + y ) + h dm − 2h x dm (1.21)

The last integral is just the x coordinate of the centre of mass, which is zero by our choice of
coordinate system. We get finally
I2 = ICM + M h2 (1.22)
8 CHAPTER 1. ROTATIONS

y



  
  
  
  
  
  
  
  
  
  
  
  
  
  
         
dm

  
  
  
  
  
  
  
  
  
  
  
  
  
  
       

  
  
  
  
  
  
  
  
  
  
  
  
  
  




  
  
   
  
   
  
   
  
 
r
 
  
   
  
   
  
          

  
 
  
 
  
 
  
 
  
 
  
 
  
  y
   
r_2





 
  
  

 
  
  

 
  
  

 
  
  

 
  
  

 
  
  

 
  
  
          
CM 
   
    
    
    
       
  
  
  2

x h−x x

Figure 1.5: Geometry for the parallel-axis theorem

which is what we set out to prove.

1.4 Newton’s 2nd. law for rotation


For a particle, or the centre of mass of a system, Newton’s 2nd law holds: If a net force F~ is applied
to the object, it acquires an acceleration ~a = F~ /m, where m, the object’s mass, is a measure of its
“inertia”, that is, its resistance to being accelerated.
In rotation, we have learned that a body may acquire an angular acceleration α as a result of
having a torque τ applied. We have also seen that the role of rotational inertia is played by I, the
moment of inertia of the body. Replacing F~ , m, and ~a with their rotational counterparts, we get
Newton’s 2nd. law for rotation:
τ = Iα (1.23)

1.5 Rolling motion


A particular case of rotation that is of some interest is rolling motion. Consider a wheel, rolling
without slipping on a flat surface. As Figure 1.6 shows, when the wheel turns through an angle dθ,
the centre of mass moves a distance equal to the arc length ds = Rdθ parallel to the surface. The
1.6. CONSERVATION OF ENERGY, INCLUDING ROTATION 9

dr

ds

Figure 1.6: Rolling

speed of the centre of mass is


ds dθ
vCM = =R = ωR (1.24)
dt dt
If the centre of mass is accelerating,

aCM = R = Rα (1.25)
dt

1.6 Conservation of energy, including rotation


In general, the motion of a body may be separated into

• the translational motion of the centre of mass with velocity ~v , and

• a rotation with angular velocity ω about an axis through the centre of mass.

Each of these contributes to the body’s total kinetic energy:


1 2 1
K = Ktr + Krot = mvCM + Iω 2 (1.26)
2 2
The work-kinetic energy theorem holds, with the full kinetic energy. In particular, if only conser-
vative forces act on a body, the total energy is conserved:

U + Ktr + Krot = constant (1.27)


10 CHAPTER 1. ROTATIONS

Exercise 1.6.1 Two cylinders of equal mass M and radius R roll without slipping down an inclined
plane, dropping through a height h. Cylinder 1 is solid, but cylinder 2 is a hollow shell. Which
cylinder wins the race?

1.7 General definition of torque


Consider a point mass m as shown in Figure 1.7. Its position relative to the point O which we have
chosen as the origin of the coordinate system is ~r, which lies in the xy plane. A force F~ is acting
on m, also in the xy plane. According to our established definition of torque, we would write the

F
φ


x
O

Figure 1.7: Calculating the torque on a point mass

torque on m about an axis that passes through O as

τ = +rF sin φ (1.28)

where the + sign is due to the fact that F~ accelerates m in a counterclockwise direction.
Compare this with the vector formed by the cross-product of ~r and F~ (see Appendix B):

~r × F~ = rF sin φ k̂ (1.29)

What we have called τ is the z component of ~r × F~ .


1.8. ANGULAR MOMENTUM 11

This suggests that we should extend our definition of torque. We shall say that the torque about
a point O on a mass at a position ~r relative to O by a force F~ is

~τ = ~r × F~ (1.30)

Some observations about this definition of torque:

• Torque is defined relative to a point rather than a rotation axis. In general, the orientation
of the rotation axis may not stay constant.

• Torque is a vector. If we limit ourselves to a rotation about a fixed axis, then the only choices
for the direction of ~τ are “up” or “down”; for that we had the sign convention mentioned
earlier.

• The torque is zero if ~r k F~ . Interestingly, this is the case of the gravitational pull by the
Sun on the planets: If we consider the torque due to gravity about the Sun, it must be zero
because the force acts along the line joining the Sun and the planet, which is also the direction
of the planet’s position vector relative to the Sun.

Exercise 1.7.1 A point mass M is at ~r = (ı̂ + ̂ + 2k̂) m from point O. Two forces act on M :

F~1 = (2ı̂ + ̂) N


F~2 = (̂ + k̂) N

Calculate the net torque on M about O.

1.8 Angular momentum


Pursuing the analogy with translational motion, we know that Newton’s 2nd law may be written
in terms of the momentum of a particle as

d~p
F~ = (1.31)
dt

where F~ is the total external force on the system, and p~ is the particle’s momentum. We may
conjecture that there is a similar quantity for rotations, and write

~
dL
~τ = (1.32)
dt

where L~ is known as the angular momentum of the particle. (Immediately after this, we will
discuss the angular momentum of a system of particles).
12 CHAPTER 1. ROTATIONS

We will show that the angular momentum is linked to the linear momentum; in fact,
~ = ~r × p~
L (1.33)
~ depends on where the origin
where ~r is the particle’s position vector. (Notice that the value of L
is taken). To see this, take the derivative with respect to time of equation (1.33):
~
dL d~r d~p
= × p~ + ~r × (1.34)
dt dt dt
The first term vanishes, since
d~r
= ~v k p~ (1.35)
dt
and in the second term we recognize d~p/dt = F~ , so
~
dL
= ~r × F~ = ~τ (1.36)
dt

1.8.1 Extension to systems of particles


The angular momentum of a system of particles is simply the sum of the angular momenta of each
particle.
In particular, the system may be a rigid body. We will limit ourselves once again to rotations
about a fixed axis, so that the body has a moment of inertia I about that axis. By analogy with
the linear momentum, the angular momentum of the body is

L = Iω (1.37)
~ only has a component along the rotation axis,
where we have dropped the vector arrows since L
but the usual sign convention still holds.

1.9 Conservation of angular momentum


From Newton’s 2nd. law for rotation, we find immediately that if there is no net external torque
~
on a system of particles, dL/dt = 0, and the angular momentum is conserved.
For example, consider a collapsing star. Stars like our Sun end their lives as “white dwarves”,
essentially the core of the star laid bare. More massive stars can continue to collapse right down
to the density of nuclear matter. A neutron star with 1.4 times the mass of the Sun may have
a radius of the order of 10 km! (See for example http://www.astro.umd.edu/ miller/nstar.html).
Since there are no external torques on the star as it collapses, it follows that
Ii If
Ii ωi = If ωf ⇒ = (1.38)
Ti Tf
1.9. CONSERVATION OF ANGULAR MOMENTUM 13

where Ti,f are the initial and final periods of revolution of the star. The moment of inertia of
a sphere about an axis through its centre scales with the square of its radius. Let us assume
conservatively that the initial radius of our doomed star is of the same order as the Sun’s, 109 m,
and Ti is also similar to the Sun’s, about 2 × 106 s. If the final radius is 104 m,

Rf2
Tf = Ti = 10−10 × 2 × 106 = 2 × 10−4 s (1.39)
Ri2

See pp 261-263 in the textbook for many more interesting applications of the conservation of angular
momentum.
14 CHAPTER 1. ROTATIONS
Chapter 2

Oscillations

15
16 CHAPTER 2. OSCILLATIONS

2.1 Oscillators
An oscillator is a physical system where the physical variables that describe it are periodic func-
tions of time. There are well-known instance of mechanical oscillators, of course, like the pendulum,
or a loaded spring; there are also electrical circuits that produce an output voltage or a current
that are periodic functions of time, and qualify as oscillators equally well. All oscillators may be
described formally in the same way. So even though we will often refer to pendulums or springs as
illustrations, we should not lose sight of the generality of the properties of oscillating systems that
we shall study here.

2.1.1 Simple harmonic motion


Position as a function of time
A simple harmonic oscillator is one where the variable of interest varies as a simple trigonometric
function. For example, consider a mass on the end of a massless spring, as shown in Figure 2.1.
Here, x is the position of the mass on the end of the spring, measured from its equilibrium position
when the mass is not oscillating1 .
If the mass is pulled down to an initial position x0 and released, it will oscillate, so that its
position as a function of time is given by

x = x0 cos ωt (2.1)

The argument of the cosine, ωt, is an angle, expressed as usual in radians; ω (Gr.: “omega”) is
called the angular frequency of the oscillator, and it is measured in rad/s .
Of course, we could start measuring t at any instant, so that the most general form of x(t) is

x = x0 cos (ωt + φ0 ) (2.2)

A little jargon:

• x0 is the amplitude of the oscillation. Since the cosine function oscillates between ±1, the
extreme values of x are ±x0 .

• The argument of the cosine function, ωt + φ0 , is called the phase. Since φ0 is the phase at
t = 0, it is called the initial phase. In our initial example, φ0 = 0.

Exercise 2.1.1 Suppose the mass was pulled down and released from rest, but we didn’t start
measuring time until a while later, when it passed through x = 0. What would be the initial phase
of this oscillation?
1
Strictly, we should be considering the effect of gravity as well. In fact, as we shall discuss later, it may be ignored,
as the only effect is to shift the equilibrium position downwards. We will show this later.
2.1. OSCILLATORS

 
 
  17


 

x = 0 : Equilibrium

Figure 2.1: A mass on the end of a spring.

It is important to note that the choice of a cosine function to describe the oscillation is arbitrary;
we could just as well have used a sine function.

Exercise 2.1.2 Write the position of the mass as a function of time in the conditions of Exercise
2.1.1, but using a sine function instead of a cosine.

Velocity and acceleration as functions of time


Once the position as a function of time has been established, the velocity and acceleration are easily
found by taking time derivatives (we may set φ0 = 0 without loss of generality):

vx (t) = −ωx0 sin ωt (2.3)


2
ax (t) = −ω x0 cos ωt (2.4)

To get a feel for the behaviour of an oscillator, you may want to look at the simulation at
http://www.phas.ucalgary.ca/physlets/shm.htm.
18 CHAPTER 2. OSCILLATIONS

2.1.2 Hooke’s law


In 1660, Robert Hooke discovered his law of elasticity. With our notation, it stated that the elastic
force applied to the mass at the end of a spring is given by

Fx = −kx (2.5)

where k is the spring constant, or elastic constant of the spring. It is a coefficient that must
determined experimentally, and it depends on characteristics of the particular spring used, such as
the material it’s made of, its shape and size.

Exercise 2.1.3 Sketch a plot of Fx (x). What is the effect of varying the value of k? Why do you
think a spring with a large value of k is called a “stiff ” spring, and one with a small value of k is
called a “soft” spring?

Since we postulated the form of the position of the mass as a function of time, it is nice to confirm
that it is consistent with Hooke’s empirical law:

Fx = max = −mω 2 x0 cos ωt = −mω 2 x (2.6)

so that we may identify r


k
ω= (2.7)
m

2.1.3 Period and frequency of the oscillation


The period T of the oscillation is the time taken to complete one full cycle, that is for the phase
to go from φ to φ + 2π. That is,

T = (2.8)
ω
The frequency f is the number of cycles per second. A moment’s thought will convince the reader
that
1
f= (2.9)
T
The units of f are inverse seconds, also known as Hertz (Hz).

Exercise 2.1.4 Find an expression for the frequency of the loaded spring, in terms of k and m.

2.2 The ideal pendulum


An ideal pendulum is simply a mass of negligible size (so that its moment of inertia in any
direction through its centre of mass is zero), suspended from a string of negligible mass. The
restoring force that plays the role of the elastic force by a spring is the component of gravity
2.2. THE IDEAL PENDULUM 19

θ
L

F_g

Figure 2.2: An ideal pendulum.

perpendicular to the string. As we shall see, for small amplitude oscillations, the pendulum also
executes simple harmonic motion.
Consider the angle that the pendulum forms with the vertical, θ. (See Figure 2.2). The torque
on the mass about the suspension point is

τ = −mgL sin θ (2.10)

and since its moment of inertia about the same point is simply

I = mL2 (2.11)

we may also write


τ = Iα ⇒ mL2 α = −mgL sin θ (2.12)
where α is the angular acceleration of the swinging mass. Therefore
g
α=− sin θ (2.13)
L
Now, it is an interesting fact that if θ is measured in radians, for small-amplitude oscillations
(θ << 1),
sin θ ≈ θ (2.14)
20 CHAPTER 2. OSCILLATIONS

Exercise 2.2.1 Verify this. Use your calculator to find sin(0.1), sin(0.01), and sin(0.001). In each
case, what is the percentage error incurred by approximating sin θ as θ?
So, for small-amplitude oscillations, we may approximate equation (2.13) as
g
α=− θ (2.15)
L
which we may recognize as the fundamental relation between position and acceleration for a simple
harmonic oscillator:
acceleration = −(positivec onstant) × position (2.16)
From this, we may write with confidence

θ = θ0 cos ωt (2.17)

(where again we’ll assume the initial phase is zero). The angular frequency is
g
ω= (2.18)
L
(compare with equation (2.7)).
Exercise 2.2.2 What is the pendulum’s period T in terms of g and L?

2.3 Elastic potential energy


Let’s return to the loaded spring. One conclusion we may draw from Hooke’s law is that the elastic
force only depends on the position of the mass. This implies that the elastic force is conservative,
and we may associate a potential energy with it.
To find the potential energy of the mass when it’s at a certain position x, we have to calculate
the work done by the elastic force as the mass moves from say x = 0 to x. Using equations (??)
and (??) (see Appendix C), and setting U = 0 at x = 0, we get
Z x
1
U (x) = −W (x = 0 → x) = − (−kx)dx = kx2 (2.19)
0 2

Exercise 2.3.1 If the principle of conservation of energy holds true, the total energy
1 1
E = mv 2 + kx2 (2.20)
2 2
should remain constant over time. Verify this for the mass on the end of a spring, bearing in mind
that it is doing simple harmonic motion.
Chapter 3

Waves

21
22 CHAPTER 3. WAVES

3.1 Waves
3.1.1 Definitions and terminology
A wave is a periodic perturbation in a quantity that is a function of position in space. Consider,
for example, a sound wave in air travelling in the x direction. The simplest form of wave is a
sine wave, so called because the pressure as a function of x and the time t has the form

p = p0 sin(kx − ωt) (3.1)

Here, p0 is the amplitude of the wave: The pressure oscillates between ±p0 . The argument of
the sine function, (kx − ωt), is called the phase of the wave. To understand the meaning of the
parameters k, consider the form of the wave at a specific time, which we can call t = 0 without loss
of generality. In that case,
p = p0 sin(kx) (3.2)
which is a periodic function of x alone. The spatial “period” is the wavelength λ, which is clearly
equal to 2π/k. The parameter k is called the wave number, defined as


k= (3.3)
λ
If instead of looking at a snapshot of the wave in time, we look at the wave at a fixed position, which
again we can call x = 0 without loss of generality, then we see that the pressure varies periodically
with time, with angular frequency ω. If T is the period of the oscillation at a fixed point, we have
again that

ω= (3.4)
T
A sound wave “travels”. Look at a peak in the wave, for example where the phase is equal to 2π.
The peak will occupy different positions x over time, given by
ω
kx − ωt = 2π ⇒ x = t (3.5)
k
This equation says that the position of the peak moves with velocity
ω
vx = (3.6)
k
along x. Because it is the velocity of a certain value of the phase, it is called the phase velocity.

Exercise 3.1.1 Show that


p = p0 sin(kx + ωt) (3.7)
represents a sound wave travelling in the negative x direction.
3.2. WAVES IN A STRING 23

3.1.2 Why study waves?


Many different phenomena have the same wave behaviour. The pressure in a sound wave, the shape
of a string under tension, the height of a liquid surface, the electric and magnetic fields in the case
of a light wave. In all these different contexts, the mathematics of wave motion is the same.

3.1.3 A general description of a sine wave


In all the types of waves mentioned above, there is a perturbation, which – if it takes the form
of a sine wave – may be written as

y = y0 sin(kx − ωt + φ0 ) (3.8)

where y0 is the amplitude and k and ω have the same meaning as before. There is a new parameter
in this equation, which we didn’t include before not to complicate matters unnecessarily: φ0 is
called the initial phase, which is simply the value of the phase at x = 0 and t = 0. If we are only
analyzing a single wave, it may be set equal to zero without restriction, but if we consider two or
more sine waves interacting, as we shall do later, then the difference in initial phase between them
is significant. More about this later.

3.1.4 Transverse and longitudinal waves


A sound wave is an example of a longitudinal wave. This means that the perturbation is along
the same direction as that in which the wave travels. In the case of surface waves in a liquid, for
example, the perturbation – the height of the surface – is perpendicular to the direction in which
the wave propagates. These are called transverse waves.

3.2 Waves in a string


As our first concrete application, we shall study waves in a string under tension. These are trans-
verse waves. The undisturbed string is straight; when the string is plucked, it is pulled sideways at
a point, and this sideways displacement travels along the string. A sine wave in a string has the
form of equation (3.8), with y representing the transverse displacement of points on the string.

3.2.1 The speed of propagation of a wave in a string


We shall apply the technique of dimensional analysis (see Appendix C) to find an expression for
the propagation speed of a wave in terms of relevant variables.
A wave in a string is generated by displacing a point sideways and releasing it. The restoring
force is the tension T applied to both ends of the string. The ability of the string to accelerate is
affected by its inertia, which may be measured by the density of the string. If the string is very
24 CHAPTER 3. WAVES

thin, it may be represented by a one-dimensional line, so the relevant density is the linear density,
or mass per unit length. It is usually represented by the Greek letter µ.
Let us settle on these two factors, then, as affecting the speed v, and write

v = CT a µb (3.9)

where C is a dimensionless constant and a and b are exponents to be determined.


Substituting SI units in this equation,

ms−1 = (kgms−2 )a (kgm−1 )b = kg a+b ma−b s−2a (3.10)

Both sides of the equation have the same units if

a+b = 0 (3.11)
a−b = 1 (3.12)
2a = 1 (3.13)

from which we find a = 1/2, b = −1/2. Therefore


s
T
v=C (3.14)
µ

Experimentally we find that C = 1 and the speed of a wave in a string is


s
T
v= (3.15)
µ

3.2.2 Energy transport by a wave


Continuing with the example of a wave on a string, although each element of the string only moves
up and down, with its height as a function of time given by

y = y0 sin(kx − ωt) (3.16)

the wave excites new parts of the string as it travels. This means that the wave is transporting
energy along the string. At what rate is this energy transported?
The kinetic energy of a certain element of the string of mass dm = µdx is
1
dK = u2 µdx (3.17)
2
where u is the speed of the oscillating element:
∂y
u= = −ωy0 cos(kx − ωt) (3.18)
∂t
3.3. THE SUPERPOSITION PRINCIPLE 25

Consequently,
1
dK = ω 2 y02 cos2 (kx − ωt)µdx (3.19)
2
We find that the rate at which kinetic energy is transmitted by the wave as
dK 1
= µvω 2 y02 cos2 (kx − ωt) (3.20)
dt 2
where we have replaced the speed of the wave v = dx/dt.
This is an oscillating function of x and t. The average kinetic energy transmitted is
 
dK 1
= µvω 2 hcos2 (kx − ωt)i (3.21)
dt 2
where h. . .i denotes an average. In fact,
1
hcos2 (kx − ωt)i = (3.22)
2
(see Appendix D for the derivation of this result). Therefore the average kinetic energy transmitted
is  
dK 1
= µvω 2 (3.23)
dt 4
This not all the energy that is transmitted by the wave, though. As each string element is
perturbed, its elastic potential energy changes too, so potential energy is transmitted with the
wave as well as kinetic energy. We will state without proof (see, however, Exercise ?? below) that
the average rate at which the wave carries potential energy is the same as that of kinetic energy:
   
dU dK
= (3.24)
dt dt
so that the average power, the rate at which energy is transmitted by the wave is
 
dK 1
hP i = 2 = µvω 2 y02 (3.25)
dt 2
A final note: Although we have derived this result for a wave in a string, it is a general property
of waves that they transport energy at a rate proportional to the square of the amplitude, and the
square of the frequency.

3.3 The superposition principle


So far we have only been concerned with waves that have a simple sine form, with a well-defined
frequency and a constant amplitude. In reality, of course, waves usually have far more complex
forms. However, a complex wave can be broken down into a sum of simple sine waves, each one
propagating independently of the others. This is the superposition principle. Thanks to it, the
analysis of wave behaviour is greatly simplified.
26 CHAPTER 3. WAVES

3.4 Interference
One typical wave phenomenon is neatly explained by the superposition principle: Interference.
Suppose that two waves, identical in every respect save for a constant phase difference, propagate
along a string. What is the resultant wave?
Call the two waves y1 and y2 , with

y1 = A sin(kx − ωt) (3.26)


y2 = A sin(kx − ωt + φ) (3.27)

The resultant wave is

y = y1 + y2 = A[sin(kx − ωt) + A sin(kx − ωt + φ)] (3.28)

We use a standard trigonometric formula for the sum of two sine functions
   
α+β α−β
sin α + sin β = 2 sin cos (3.29)
2 2

to write y as
y = 2A cos(φ/2) sin(kx − ωt + φ/2) (3.30)

This is again a sine wave, but its amplitude depends on the phase difference φ between the two
waves. If φ = 0 so that the two waves coincide, then y is twice each of the two component waves.
This situation is called constructive interference. At the other extreme, if φ = π so that the
two waves are half a cycle out of phase, y = 0: The waves cancel! This is called destructive
interference.

3.5 Standing waves


Consider now two identical waves travelling in opposite directions along a string:

y1 = A sin(kx − ωt) (3.31)


y2 = A sin(kx + ωt) (3.32)

Exercise 3.5.1 Consider a “peak” in each wave, where the phase is equal to zero. What is x(t),
the position of the peak as a function of time? From this, what is the direction of propagation of y1
and y2 ?
3.6. RESONANT MODES OF A STRING 27

To find the resultant wave, use equation (3.29) again:

y = 2A sin(kx) cos(ωt) (3.33)

This is an oscillating function of x – sin(kx) – with an amplitude that is a periodic function of


time – 2A cos(ωt) –. This is not a propagating wave, but rather a stationary wave whose amplitude
increases and diminishes periodically: A standing wave.
Standing waves in a string are easily generated by tying down one end of the string and shaking
the other end. A wave travels down the string and is reflected at the fixed end (Newton’s 3rd. law:
The string pulls on the wall where the string is tied, and the wall pulls back, shaking the string
and sending a wave in the opposite direction!). Very quickly the reflected wave meets the incoming
wave, and the two interfere to produce a standing wave.

3.6 Resonant modes of a string


If both ends are tied down and the string is plucked, waves travel in both directions, and are
reflected in both directions. Under certain conditions, all these waves will interfere to produce a
standing wave. Say the length of the string is L. If a standing wave of the form

y = ymax sin(kx)cos(ωt) (3.34)

is produced, y must be zero at x = 0 and at x = L at all times. Therefore

sin(kL) = 0 ⇒ kL = nπ (3.35)

where n = 1, 2, . . . . Remembering that k = 2π/λ, this means that


2L
λ= (3.36)
n
Each of these possible standing waves is called a resonant mode of the string. The lowest frequency
mode, with n = 1, is called the fundamental mode or first harmonic1 . Higher frequency modes
are called second, third, etc. harmonics.

3.7 Intensity of sound and sound level


3.7.1 The inverse square law
The energy emitted by a source of sound is carried away by the waves, as we have seen. If the
source is point-like, and the energy is spread uniformly in all directions, the wave fronts are spherical,
1
The nomenclature differs between authors. Some call the first harmonic the mode above the fundamental. Here,
however, we shall follow the same nomenclature as in the textbook.
28 CHAPTER 3. WAVES

centered at the source. The surface of such a wave front is proportional to the square of its radius:
A = 4πr2 . If a detector of collecting area a is located at a distance r from the source, the fraction
of energy that the detector collects is a/A, which is proportional to 1/r2 . Thus, the power (energy
per second) collected by the detector is inversely proportional to its distance from the source.

3.7.2 Intensity
Since the power received by a detector is also proportional to its area, it is useful to have a measure
of power per unit area at the detector. This is the intensity of the sound, and it is measured
in W/m2 . If the power emitted by a point source is P , the intensity received by a detector at a
distance r from the source is
P
I= (3.37)
4πr2

3.7.3 The decibel scale


The dynamic range of human hearing is surprisingly large: From about 10−12 W/m2 to about
1 W/m2 at the threshold of pain. Therefore a convenient measure of intensity is a logarithmic
one.
The sound level of a signal β is defined as

I
β = (10 dB) log10 (3.38)
I0

where dB stands for decibel. I0 is arbitrarily chosen to be 10−12 W/m2 , approximately the lower
limit of human hearing. Thus the dynamic range of human hearing goes from approximately 0 dB
to 120 dB.

3.7.4 Hearing curves


In fact, the sensitivity of the human ear depends on frequency: Two sounds at different frequencies
may be perceived as being equally loud, although their sound level is quite different. Equal loud-
ness curves may be drawn (see http://hyperphysics.phy-astr.gsu.edu/hbase/sound/eqloud.html),
by comparing the perceived loudness of a sound at a certain frequency with the sound level in dB
at 1000 Hz. For example a sound of 120 dB at 20 Hz (about the lowest frequency that we can hear)
is equally loud to a sound of only 90 dB at 1000 Hz. As a general rule, we may say that at the
frequency extremes of hearing the sensitivity drops. But it’s also interesting to observe that we are
exceptionally sensitive to frequencies around 3 - 4 kHz. This coincides with a resonant frequency of
the auditory canal (see http://hyperphysics.phy-astr.gsu.edu/hbase/sound/maxsens.html). It has
been suggested that the evolutionary significance of this resonance is that this frequency coincides
with a baby’s cry!
3.8. THE DOPPLER EFFECT 29

3.8 The Doppler effect


When either the source of sound, or the observer, or both, are moving relative to the medium, the
perceived frequency is not the same as the frequency emitted by the source. This phenomenon is
called the Doppler effect, after its discoverer Christian Doppler. (See http://www-groups.dcs.st-
and.ac.uk/ history/Mathematicians/Doppler.html). Because of the precise relation between the
emitted frequency and the detected frequency, it is possible to infer the relative velocity of source
and observer from measurements of the frequency shift. Although we shall only deal here with the
classical Doppler effect, which applies to sound waves, the Doppler effect also applies to light waves
in a modified form. It is this application that has yielded extremely useful results in astrophysics,
notably determining the expansion of the universe.
We will consider separately the cases of a moving detector and a moving source.

3.8.1 Moving detector, stationary source


The stationary source S emits waves with freqency f , and these travel towards the detector D with
speed v. Figure 3.1 is a time-position diagram (notice the unusual choice of axes), showing the

x_D=v_D t
t

x_2 = v ( t − T )

T’ x_1 = v t

0 x

Figure 3.1: Moving detector, stationary source.

lines corresponding to two successive wave fronts,

x1 = vt (3.39)
x2 = v(t − T ) (3.40)
30 CHAPTER 3. WAVES

where T = 1/f is the period of the signal. The diagram also shows the detector, whose position as
a function of time is
xD = vD t (3.41)
For simplicity, we have chosen the detector to coincide with the source at t = 0, precisely when the
first wave front is emitted.
The second wave front reaches the detector when x2 = xD , that is, at a time T 0 given by

v(T 0 − T ) = vD T 0 (3.42)

so that
v
T0 = T (3.43)
v − vD
and the frequency measured by the detector is therefore
v − vD
f0 = f (3.44)
v

Exercise 3.8.1 Redo the calculation above for the case that the detector is moving towards the
source with speed vD , and verify that in this case
v + vD
f0 = f (3.45)
v

3.8.2 Moving source, stationary detector


Figure 3.2 shows the time-position diagram for this situation.
Now the source’s position as a function of time is

xS = vS t (3.46)

where once again for simplicity we let the source coincide with the detector at the time when the
first wave front is emitted. Then when the second wave front is emitted at time T , the source
has reached the position vS T , and the second wave front moves away from the source towards the
detector with speed v, so that its position as a function of time is

x2 = vS T − v(t − T ) (3.47)

This wave front will reach the stationary detector at a time T 0 when x2 = 0, that is,

vS T − v(T 0 − T ) = 0 (3.48)

which gives
v + vS
T0 = T (3.49)
v
3.8. THE DOPPLER EFFECT 31

t x_S = v_S t

T’
x_2 = v_S T − v( t − T )

0 x
v_S T

Figure 3.2: Moving source, stationary detector.

and
v
f0 = f (3.50)
v + vS

Exercise 3.8.2 Like before, verify that if the source is moving towards the detector,
v
f0 = f (3.51)
v − vS

3.8.3 The general Doppler effect


If both the source and the detector are moving relative to the medium, the frequency measured by
the detector is the result of combining the two previous results:
v ± vD
f0 = f (3.52)
v ± vS
The correct signs are determined by bearing in mind the following simple rule (quoted here verbatim
from HRW):
When the motion of detector or source is toward the other, the sign on its speed must
give an upward shift in frequency. When the motion of detector or source is away from
the other, the sign on its speed must give a downward shift in frequency.
32 CHAPTER 3. WAVES
Chapter 4

Fluids

33
34 CHAPTER 4. FLUIDS

4.1 Introduction
4.1.1 Solids, liquids, and gases

Fluids – a term that includes both liquids and gases – can flow and change shape, unlike solids.
This difference in basic properties has a simple microscopic explanation. As Figure 4.1 illustrates,
atoms in solids are very closely bound in a rigid lattice; in a liquid, the atoms are a little further
apart, and the bonds between them are weaker, so they can slide past each other; in a gas, the
atoms are so far removed from each other that they hardly interact at all except when two happen
to collide. It follows that in a given volume, atoms will be most closely packed in a solid, a little
less so in a liquid, and far less in a gas. The typical densities of materials in the three states are
consistent with this picture. Typical densities of liquids and solids are on the order of 103 kg/m3 ,
while gases at STP1 have typical densities on the order of 1 kg/m3 .

4.1.2 Density and pressure

We shall be looking for the basic laws of motion that govern fluids, analogous to Newton’s laws
in particle mechanics. However, we can no longer apply particle concepts like mass or force to a
continuous medium. Rather we speak of the density and the pressure at a point in a fluid.
The density is the amount of mass per unit volume at a “point” in the fluid. Since a point has
no volume, this may seem puzzling. What we really mean is a very small volume element of the
fluid, much smaller than the size of the fluid sample, but large enough to contain a large number
of atoms. If the fluid is uniform, the density is the same at all points, so even a large volume of
the fluid has the same density.
In a similar spirit we define the pressure at a point in the fluid as the magnitude of the force
per unit area on a tiny disk at that point. Interestingly, pressure is a scalar: The orientation of the
disk is irrelevant.
In the SI system of units, density is measured in kg/m3 , and pressure is measured in N/m2 , a unit
that also goes by the name of Pascal (Pa), after Blaise Pascal, philosopher and mathematician. (See
http://www-groups.dcs.st-and.ac.uk/ history/Mathematicians/Pascal.html for a short biography of
this very interesting individual).

4.2 Fluid statics


We begin our study of fluids by considering fluids that are in equilibrium, and therefore static.
1
STP means “Standard Temperature and Pressure”, defined as 0 ◦ C and 1 atmosphere of pressure (see section ).
These are conditions typical of the Earth’s atmosphere at sea-level.
4.2. FLUID STATICS 35

4.2.1 Variation of pressure with depth


Consider first an incompressible fluid, typically a liquid. By definition, an incompressible fluid
has uniform density ρ. Isolate a column in the fluid of cross-sectional area A extending from a
depth h0 to a depth h1 (see Figure 4.2).
The pressure on the lower surface must be greater than the pressure on the top surface, to
balance the weight of the fluid in the column. Newton’s 2nd. law in the vertical direction gives
ρA(h1 − h0 )g = p1 A − p0 A ⇒ p1 = p0 + ρg(h1 − h0 ) (4.1)
Thus the pressure increases linearly with depth in an incompressible fluid.
Exercise 4.2.1 If the pressure on the surface of a body of water is 105 Pa, a typical value of the
atmospheric pressure, and the density of water is 103 kg/m3 , how far down would the pressure be
twice the atmospheric pressure at the surface?
We get a different result when we consider a compressible fluid, such as the atmosphere.
Since the density can now vary with height, we must consider a thin slice of the fluid at a height z
where the density ρ(z) can be assumed to be uniform. See Figure 4.3.
The balance of forces in the vertical direction gives
−ρ(z)Agdz = pA − A(p + dp) ⇒ dp = −ρ(z)gdz (4.2)
In order to calculate the profile of p versus height z, we would have to know the form of ρ(z).
For example if we assume the gas follows the ideal gas law, then2
ρ
p = kT (4.3)
m
where m is the mass of an atom, T is the temperature of the gas measured in degrees Kelvin (K),
and k is Boltzmann’s constant, k = 1.38 × 10−23 J/K. Substituting ρ in equation (4.2) we get
dp mg
=− dz (4.4)
p kT
Notice that kT /mg has dimensions of length. This quantity is called the scale height H. To find
the pressure at any height z, we integrate between a base height z = 0 where the pressure is p0 ,
and z: Z p
1 z 0
Z
dp
=− dz (4.5)
p0 p H 0
which gives 
p z
ln = − ⇒ p = p0 e−z/H (4.6)
p0 H
The pressure decreases with height, not linearly but exponentially. The density, being proportional
to the pressure, also decreases linearly with height. This makes sense: The lower layers of the fluid
should be more compressed by the weight of the fluid on top, and therefore more dense.
2
This statement is offered without proof, purely for the purpose of illustration
36 CHAPTER 4. FLUIDS

4.2.2 Buoyancy and Archimedes’ Principle

It is a common experience that objects sink or float in a fluid, depending on their density. Legend
has it that Archimedes of Syracuse discovered the exact mechanism for this. (See
http://www-groups.dcs.st-and.ac.uk/ history/Mathematicians/Archimedes.html)
Consider for example the fearsome beast in Figure 4.4 (Credit: D.J.,
http://www.clay.k12.fl.us/wec/Animal.htm). If the animal was removed, it would be replaced by
a portion of fluid of the same shape and volume which would be at rest, sustained by a force from
the fluid equal to its weight. Since the fluid is unaware that this space is now occupied by a shark,
it applies the same force to the shark. This is called the buoyancy force.
To put it mathematically, call ρ the average density of the shark, ρf the density of the fluid, and
V the volume of the shark, which is the same as that of the displaced fluid. The shark’s weight is
ρV g, and the buoyancy force, equal to the weight of the displaced fluid, is ρf V g. The net (upward)
force on the shark is then
(ρ − ρf )V g (4.7)

which is positive if ρ < ρf , in which case the shark floats upward, and negative if ρ > ρf , in which
case the shark sinks. If ρ = ρf , the shark will neither sink nor swim; it is said to experience neutral
buoyancy.

Exercise 4.2.2 What is the weight of the shark under water? (Assume the shark is more dense
than the surrounding water, which is hardly surprising). If you were able to measure the shark’s
weight outside the water, as well as under water, how would you calculate its density relative to
water from that information?

4.3 Fluid dynamics


4.3.1 The velocity field in a fluid

We turn now to the problem of a moving fluid. Our objective is to find conservation principles
analogous to the known principles that apply to particle mechanics. To describe the motion of
the fluid, we need to keep track of the velocity of the fluid at each point. Notice that this does
not mean to follow a particular fluid element as it moves and record its velocity. It means that
we focus on a fixed point in the space occupied by the fluid, and record the velocity of each fluid
element that passes through that point. A map of the velocities at all points in the fluid is called
the velocity field. It can be visualized approximately by drawing velocity vectors on a grid. Here
are some good examples: http://virga.sfsu.edu/crws/jetstream.html.
If the velocity field does not vary with time, the flow is said to be steady, or stationary.
4.3. FLUID DYNAMICS 37

4.3.2 Streamlines
If you looked at some of the wind velocity maps mentioned above, you could see that the arrows
could be quite easily joined by smooth curves. Such curves are called streamlines, defined as
curves which are tangent to the velocity vectors at each point. If the flow can be visualized using
streamlines, it means that any fluid element that finds itself at some point between two given
streamlines, remains between those streamlines at all future times. This is because of the definition
of streamlines: Since they are tangent to the velocity of the fluid at all points, it follows that no
fluid element can cross a field line. The flow is said to be laminar, since the fluid moves in layers.
The opposite of laminar flow is turbulent flow. There, streamlines are not defined. We shall
restrict our analysis to laminar flow only.

4.3.3 The continuity equation


The first conservation principle we apply is the conservation of mass. Figure 4.5 shows a portion of
a fluid flowing between two streamlines. At the left end, with cross-section area A1 , the fluid moves
with speed v1 . At the right end, with a different cross-section area A2 , the speed is also different,
v2 . Since the fluid is contained within the streamlines, the amount of mass that enters from the
left during a time dt is the same that leaves at the right during the same time. Symbolically,
ρ1 v1 A1 dt = ρ2 v2 A2 dt (4.8)
and therefore, within the same streamlines anywhere
ρvA = constant (4.9)
This is known as the continuity equation. The quantity ρvA is called the mass flow rate,
measured in kg/s.
If the fluid is incompressible, ρ = constant, and we can further say that
vA = constant (4.10)
Here, vA has dimensions of m3 /s; it’s called the volume flow rate.
In the case of an incompressible fluid, the continuity equation implies that if a fluid element is
forced through a constriction, it will speed up, or if it is allowed to widen, it will slow down. In
other words, the fluid is accelerated when the cross-section area changes, and that acceleration can
only be brought about by a pressure difference across the fluid element. The pressure difference
that produces a certain acceleration is determined by conservation of energy, to which we turn now.

4.3.4 Bernoulli’s law


We’ll consider now the principle of conservation of energy as applied to a fluid. In this section,
we shall argue something quite surprising: The pressure in a fluid behaves mathematically as a
potential energy function.
38 CHAPTER 4. FLUIDS

As we show in Appendix E, if a particle moves along a direction x under the influence of a


conservative force F~ , and the potential energy is U (x), there is a simple relation between the
potential energy and the component of the force in the direction of motion:
dU
Fx = − (4.11)
dx
Now consider instead a fluid element of width dx and cross-sectional area A moving between a pair
of streamlines. Considering the pressure difference across the fluid element, the net force in the x
direction is
pA − (p + dp)A = −Adp (4.12)
Now, the volume of the fluid element is Adx; so, interestingly, the force per unit volume along x is
dp
Fx = − (4.13)
dx
so p mimics a potential energy!3
Conservation of energy – per unit volume – then implies that
1
p + ρv 2 = constant (4.14)
2
along a streamline. Furthermore, if the fluid is also subject to a conservative force, its associated
potential energy should be included in the equation. For instance, if the fluid is moving in the
gravitational field near the surface of the Earth, along any streamline
1
p + ρgh + ρv 2 = constant (4.15)
2
This is known as Bernoulli’s equation, after the talented 18th. century mathematical physicist
Daniel Bernoulli.

3
It is important to realize that although the pressure has the mathematical properties of a potential energy, it
isn’t really a potential energy. A conservative force requires a source: The gravitational pull on the Moon is due to
the Earth (and other bodies). There is no “source” of the pressure.
4.3. FLUID DYNAMICS 39

Solid

Liquid

Gas

Figure 4.1: A microscopic view of solids, liquids, and gases.


40 CHAPTER 4. FLUIDS





 

 


p_0

Cross−sectional area A
0




 

 
 

 
 




 

 

 
   
h_0



 
 
 
 





 

 

 




 

 
 

 

Weight 




 

 

 




 

 
 

 
 




 

 

 
h_1




 
 

h

p_1

Figure 4.2: Variation of pressure with depth in an incompressible fluid.


4.3. FLUID DYNAMICS 41

p+dp Cross−sectional area A

z+dz
z

Figure 4.3: Variation of pressure with depth in a compressible fluid.

Figure 4.4: D.J.’s hammerhead shark.


42 CHAPTER 4. FLUIDS

A_1


  



   A_2


        


     
    








      
v_2


v_1


 


 


v_2 dt

v_1 dt

Figure 4.5: Conservation of mass.


Appendix A

Derivations for the final exam

43
44 APPENDIX A. DERIVATIONS FOR THE FINAL EXAM

A.1 Derivations for the final exam


The following is a list of derivations we have seen in the course, one of which will be included in
the final exam, with exactly the same wording as given here. Beside each one, a link is provided to
the page in these notes where the derivation may be found.
1. The parallel-axis theorem: See Section 1.3.4. The moment of inertia of an extended body
about a fixed axis is the integral Z
I= r2 dm

where r is the distance from the mass element dm to the axis, and the integral is taken over
the whole body.
Show that for an arbitrary fixed axis, I is related to ICM , the moment of inertia with respect
to a parallel axis that passes through the centre of mass, by
I = ICM + M h2
where M is the mass of the body, and h is the separation between the two axes.
In your proof, you may use the definition of the coordinates of the centre of mass:
Z
~rCM = ~r dm

where ~r is the position of a mass element dm, and the integral is again taken over the whole
body.
2. Torque and angular momentum: See Section 1.8. Show that for a point mass m with
~ and a net torque ~τ about a point P
angular momentum L
~
dL
~τ =
dt
3. A standing wave on a string: See Sections ?? and ??. Show that if two waves of identical
angular frequency ω and wave number k travel in opposite directions along a string, they
combine to form a standing wave of the form
y = A(t) sin(kx)
where A(t) is a time-dependent amplitude. Give an expression for A(t).
Use the following trigonometrical relation:
   
α±β α∓β
sin(α ± β) = 2 sin cos
2 2
4. Variation with depth of pressure in an incompressible fluid: See Section 4.2.1. Derive
the pressure p as a function of depth h in an incompressible fluid of density ρ.
Appendix B

The vector cross product

45
46 APPENDIX B. THE VECTOR CROSS PRODUCT

B.1 The vector cross product


B.1.1 Definition
The cross product is an operation between two vectors in 3 dimensions that gives a third vector
as a result. It is written as
~a × ~b (B.1)
and it’s pronounced “a cross b”.
The direction of the vector ~a × ~b is perpendicular to both ~a and ~b, and given by a right-hand
rule: Point the fingers of the right hand along the first vector ~a, and sweep them towards the
second vector ~b. The thumb points automatically in the direction of ~a × ~b.
The magnitude of the cross product is

|~a| × |~b| = |~a||~b| sin φ (B.2)

where φ is the (small) angle between ~a and ~b.


One important consequence of this definition is that the cross product of two vectors that are
parallel is zero, since φ = 0 in that case.

B.1.2 The cross product is non-commutative


With the cross product, the order of the factors does matter! From the definition,

~a × ~b = −~b × ~a (B.3)

B.1.3 The x, y, and z components of a cross product


If we are given

~a = ax ı̂ + ay ̂ + az k̂ (B.4)
~b = bx ı̂ + by ̂ + bz k̂ (B.5)

we can construct the cross product as a sum of terms containing the cross products of all possible
combinations of ı̂, ̂ and k̂:

~a × ~b = ax bx ı̂ × ı̂ + ax by ı̂ × ̂ + . . . (B.6)

So, we need to construct all the possible cross products of unit vectors. Actually, the only ones we
need are:

ı̂ × ̂ = k̂ (B.7)
̂ × k̂ = ı̂ (B.8)
k̂ × ı̂ = ̂ (B.9)
B.1. THE VECTOR CROSS PRODUCT 47

The cross products of each unit vector with itself are all zero, and a product like ̂ × ı̂ is obtained
as −ı̂ × ̂.
Now we can write the components of ~a × ~b. We’ll leave it as an exercise to show that

(~a × ~b)x = ay bz − az by (B.10)


(~a × ~b)y = az bx − ax bz (B.11)
(~a × ~b)z = ax by − ay bx (B.12)

Exercise B.1.1 Make up a mnemonic rule to remember the components of a cross product. (Hint:
Cyclic permutations of indices).
√ √ √ √
Exercise B.1.2 ~a = 1/ 2ı̂ + 1/ 2̂ and ~b = −1/ 2ı̂ + 1/ 2̂. Calculate ~a × ~b.
48 APPENDIX B. THE VECTOR CROSS PRODUCT
Appendix C

Dimensional analysis

49
50 APPENDIX C. DIMENSIONAL ANALYSIS

C.1 The magic of dimensional analysis


The technique of dimensional analysis may seem almost magical: It is possible to derive the
form of physical laws without knowing any physics! It’s a two-step technique: To illustrate it, we
shall use it here to derive the form of the period of an ideal pendulum.

C.1.1 Step 1: What does the quantity of interest depend on?


What can the period of a pendulum depend on? By playing with a pendulum, we can conclude
that it depends on the length of the pendulum L. We can also surmise that it depends on the
gravitational field g, even if we can’t leave the Earth to test the idea. And forgetting everything
we’ve learned about pendulums, we might imagine that it depends on the mass m of the pendulum
bob. We will, however, make the strong assumption that the period is independent of the amplitude
of the oscillation. Summarizing, we can write

T = f (L, g, m) (C.1)

C.1.2 Step 2: Narrow down the possible functions by demanding that the di-
mensions are right
The fundamental assumption that we make next is that the function f is in fact a power law:

T = CLa g b mc (C.2)

where the exponents a, b, and c are yet to be determined, and C is a dimensionless constant.
Necessarily, the dimensions of both sides of this equation must be the same. Using SI units, this
means that
s = ma (ms−2 )b kg c = ma+b s−2b kg c (C.3)
This leads to a set of equations for the exponents:

a+b = 0 (C.4)
−2b = 1 (C.5)
c = 0 (C.6)

Notice that we find straight away that the exponent of m is zero, so the period T is independent
of the mass. Solving for the remaining exponents, we get b = −1/2 and a = 1/2.
Now we can write s
L
T =C (C.7)
g
This is as far as we can go with dimensional analysis. We could find C experimentally, by
measuring the period of a pendulum and varying the length L. Since we know that the general
C.1. THE MAGIC OF DIMENSIONAL ANALYSIS 51

form of T as a function of L is given by our last result, we could plot T 2 vs. L, and from the slope
find C. We would find that it is very close to 2π.
Indeed, one of the useful features of dimensional analysis is that it suggests experiments. An-
other extremely useful application of dimensional analysis is in modelling. To simulate how a 1
meter long pendulum on the Moon might behave, we would observe that √ g on the Moon is about
1/6 the value on Earth, so we would construct a pendulum of length 6 m here at home.
52 APPENDIX C. DIMENSIONAL ANALYSIS
Appendix D

Average value of cos2(x)

53
54 APPENDIX D. AVERAGE VALUE OF COS 2 (X)

D.1 Average value of cos2 (x)


In general, the average of a function f (x) over a certain interval [a, b] is calculated as
Z b
1
hf (x)i = f (x)dx (D.1)
b−a a

Since cos2 (x) is a periodic function of x with period 2π, we only need to find the average over one
cycle, say [0, 2π]:
Z 2π
2 1
hcos (x)i = cos2 (x)dx (D.2)
2π 0
All’s fair in love, war, and integration, so we’ll use a dirty trick. A moment’s thought will
convince you that
hcos2 (x)i = hsin2 (x)i (D.3)
since sin(x) is the same function as cos(x), but with a phase shift of π, (remember, the “phase” is
the argument of a trig function), so the averages should be the same. Now,
1
cos2 (x) = 1 − sin2 (x) ⇒ hcos2 (x)i = 1 − hsin2 (x)i → hcos2 (x)i = hsin2 (x)i = (D.4)
2
Appendix E

Work and potential energy

55
56 APPENDIX E. WORK AND POTENTIAL ENERGY

E.1 Work and potential energy


In this section, we revisit briefly the familiar topics of work and potential energy, now with the
benefit of calculus.
A particle moves in the x direction under the action of a force F~ . As the particle is displace by
dx, the amount of work done on it is
dW = Fx dx (E.1)
and the total amount of work done in moving the particle from xi to xf is
Z xf
∆W = Fx dx (E.2)
xi

If F~ is a conservative force, we can associate a potential energy function U to it, defined simply as

∆U = −∆W (E.3)

In this case, we may rewrite equation (E.1) as

dU = −Fx dx (E.4)

so that there is a simple relation between the force and the potential energy:
dU
Fx = − (E.5)
dx
Here, for simplicity, we restricted the motion of the particle to one dimension. In general, the
components of the force along x, y, and z are found from the potential energy U as
∂U
Fx = (E.6)
∂x
∂U
Fy = (E.7)
∂y
∂U
Fz = (E.8)
∂z

Vous aimerez peut-être aussi