Block 9 Notes

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

Block 9 1

Block 9 Notes

Physics 206 Fall 2002


Block 9 2

Overview
We begin Block 9 by using the methods and tools we have
developed this semester to analyze the oscillatory motion we first
introduced in Physics 205 as we developed our model of matter. We
now have the tools necessary to understand not only the energy aspects
of oscillatory motion, but also the detailed relationship of the motion to
the forces involved. For example, we can now tackle questions such as,
“What determines the period of a pendulum or an oscillating mass on a
spring?” In this process we will discover that all oscillatory motion,
although arising in an infinite variety of physical systems, is very similar;
the details of all oscillatory motion can be represented in the same rather
simple way. As we study oscillations, we will make a conscious effort to
integrate the energy approach with the detailed forces approach.
We also begin our study of waves in this first block of Physics
206 and will continue in the second block. In this block we focus on
what waves are and how to describe one-dimensional waves
mathematically.
Waves in material media are very closely related to vibrations of
the particles in the media. We will see that the mathematical description
of one-dimensional waves is very similar to the description of vibrations.
We postpone until next block the fascinating topic of what happens when
two waves are in the same place at the same time. Most of the interesting
phenomena associated with sound and light (which can be modeled as a
wave motion) are due to the interactions of more than one wave.
An important message of the entire Physics 205/206 course is that
one of the distinguishing features of physics is the continual striving for
general principles and simple models that can be applied to large classes
of phenomena. In our study of wave phenomena we very consciously
take this approach; the focus is on the model and its representation, and
not on one or another of an almost unlimited number of individual
phenomena associated with sound, light, TV and radio waves, microwaves,
etc. Our goal is to enable you to develop a useful understanding of wave
behavior that you can then apply to any phenomenon that can be modeled
as a wave, whether on a quiz or the course final, or more importantly,
throughout your everyday life and in your professional or scientific
career.
There are four important goals associated with this block. The
first two have to do with oscillations, the second two with waves. With
respect to oscillations, you should become much more familiar with the
phenomenon of small oscillations and how for small amplitude, all
oscillations, regardless of how different the physical systems, exhibit the
same fundamental behavior. Second, you should become comfortable
with the mathematical representation of oscillatory motion, so that it is a
useful tool to understand the physics, rather than a hurdle to get over.
The two goals regarding waves are similar: To become familiar with wave
phenomena and how we analyze them and secondly, to understand the
mathematical representation of one-dimensional harmonic waves
sufficiently, so that we can use it as a tool in the next block to help us
understand the physics of sound and light (electromagnetic) waves.

Oscillatory Motion
Physics 206 Fall 2002
Block 9 3

Introduction
Periodic or oscillatory motion is common throughout the universe,
from the smallest to the largest distance and time scales. This kind of
motion is related to the forces acting between objects by Newton’s 2nd
law, just as all motions are. To have oscillatory motion, there must be a
restoring force that acts in a direction to cause an object to return to its
equilibrium position. A particularly common type of oscillatory motion
results when the magnitude of the restoring force is directly proportional
to the displacement of the object from equilibrium:
Frestoring = -k x.
This is the force law characteristic of a spring that is stretched or
compressed from its equilibrium position. The oscillatory motion that
results from this force law is known as simple harmonic motion (SHM).
Even when the force law is not as simple as Frestoring = -k x for
arbitrary values of x, it turns out that for an object that oscillates about an
equilibrium position, this linear law provides an accurate description for
small oscillations. Thus, we can make a very strong statement: essentially
every system that vibrates, does so in SHM for small amplitudes of
vibration.
Because it is so common, it is worth spending some effort under–
standing SHM and the different ways to represent it. Another reason for
focusing on SHM is that periodic wave motion is the interconnected
vibrations of many, many oscillators, each vibrating in SHM.
Simple Harmonic Motion
Our approach is to use the tools we have at our disposal, namely,
Newton’s 2nd law, to analyze the motion of several different physical
systems that exhibit oscillatory motion. We will look for common
features of the motion and its description. Then, we will generalize the
description and representation. In this process we will develop explicit
mathematical expressions to represent the motion and we will see how
properties of the motion such as the period of oscillation are related to the
physical parameters of the particular phenomena. In this process we will
revisit the energies involved in oscillating systems and gain a deeper
understanding of the energy relationships.
First system: mass on a spring
We consider a mass hanging on a spring. There are two forces
acting on the mass: the pull upward of the spring and the gravity force of
the Earth pulling down. We saw previously that if we take x to be the
distance from the equilibrium position of the mass as it hangs motionless
on the spring, then the net force has the form
∑F
= -kx
Now we apply Newton’s 2nd law: net force equals mass times
acceleration
-k x = m a, or
d2x
-k x = m
dt 2

Physics 206 Fall 2002


Block 9 4

This is called a differential equation, because it involves derivatives of x.


A standard way to write this equation that will be useful as we compare
different systems is
d2 k
a= 2
x(t) = - x(t)
dt m
Let’s note several things about this equation. Its solution will be a
mathematical expression that gives the position x as a function of the time
t. The equation says that if we differentiate this function twice, we get
back the same function multiplied by the negative constant -k/m
d2
(k and m are both positive constants). Also, the acceleration, a = 2 x(t) ,
dt
is not constant. Rather, the acceleration is proportional to the
displacement from equilibrium, but with the opposite sign.
What kind of function, when differentiated twice, gives back the
same function, but with a negative constant coefficient? Perhaps you
remember from your calculus course which function has this property. If
you don’t, that is OK. What’s important is to understand the properties
of the solution, not how to get the solution. There are two functions that
have the property we desire. One is the sine function and the other is the
cosine function. The second derivative of Asin(b t) with respect to t
(when A and b are constants) is -b2Asin(bt). That is,
d2
2
Asin bt = −b 2 sin bt
dt
And similarly for the cosine function.
We compare this equation to the equation for the mass on a spring,
d2 k
2
x(t) = - x(t)
dt m
we notice that they are the same if b2 equals k/m. If we
make that substitution, we get

k
x(t) = Asin( t)
m
Try differentiating twice with respect to t and see if you don’t get the
function x(t) back multiplied by the negative constant -k/m.
These two functions (the sine and the cosine) are solutions of the
differential equation we obtained by applying Newton’s 2nd law to the
mass hanging on the spring. These functions repeat every time the angle
bt increases by 2π. Thus, the time to complete one oscillation is that value
of t that satisfies the relation bt = 2π. This time is called the period and is
denoted by the letter T. It is equal to 2π/b.
2π m
T= = 2π
b k
Note that we know the period if we know the values of the factors that
appear in Newton’s 2nd law (mass and spring constant).

Physics 206 Fall 2002


Block 9 5

We can now write the differential equation in terms of T,


d2 2π 
2
x(t) =-  x(t)
dt 2  T 
and the possible solutions also in terms of T:
2π t
x(t) = Asin
T
or
2π t
x(t) = Acos
T

Before pursuing the analysis of the spring mass system further,


we will look at another system. Then we will generalize our results and
discuss them in much more detail.
Second system: simple pendulum
We consider a mass hanging on a lightweight string. The mass
swings back and forth when pulled aside and released. How do we apply
Newton’s 2nd law?
θ l
We first identify the objects and all the forces acting on the objects.
Then, the Net Force acting on any particular object is equal to the
product of the mass and acceleration of that object. In the case of our m
pendulum, the object of interest is the bob. (In our model, the mass of
the string is negligible.) Two forces act on the bob - the tension in the
string directed along the string, and the gravitational pull of the earth
straight down on the bob. The vector sum of these two forces is the net
force or unbalanced force. The motion is constrained to be along the arc
of a circle with radius equal to the length of the string, l. The tangential
component of the net force, that is, the force tangent to the path the bob
takes, is the component that causes the bob to speed up or slow down
along this path. (The component of the net force along the string causes
the bob to move in a circle, and is dependent on the instantaneous speed.
We don’t need to be concerned with this radial force now.)
To proceed, we draw a force diagram, showing
the forces acting on the bob.
Applying Newton’s 2nd law along the tangential
direction gives θ
m g sinθ = -m atangential T T
The minus sign tells us that atangential is opposite to
the direction of increasing θ. It is useful to express
atangential in terms of θ. Since atangential is the second m
derivative of a distance moved along the arc, and since
a distance along the arc is simply the product l θ, mgcos θ
atangential = l d2θ/dt2. Then mgsin θ
wt
= mg
d θ 2
m g sinθ = -ml ,
dt 2
Physics 206 Fall 2002
Block 9 6

and canceling out the mass


d 2θ
g sin θ = -l .
dt 2
This looks almost like our equation of motion for a mass on a spring The
difference is we have a sinθ instead of a θ on the left hand side. Perhaps
for small oscillations, that is, small values of θ, we can replace sinθ with θ.
Does sinθ = θ for small values of θ ? Try it with your calculator. You
should find that θ measured in radians is a good approximation for sinθ
up to 1/10 or even 1/5 radian (5˚ to 10˚).
If we make this approximation (substituting θ for sinθ and then
group the constants together on the left hand side) we get:
g d 2θ
- θ= 2
l dt
If we put this in standard form, we can easily compare it to the
equation we got for the mass oscillating on a spring.
d 2θ g
simple pendulum: a= 2
= - θ(t)
dt l

d2x k
mass on spring: a= 2
= - x(t)
dt m

Notice the similarity in these two equations. Except for the name of the
variable, θ or x, which is arbitrary, they have the identical form.
We saw before, that in terms of the period to make a complete
oscillation, we could write the expression for x(t) as:

 2π 
2
d2
2
x(t) = -   x(t) , where
dt T
 2π  = k
2
m
→ T = 2π
 T  m k
Now by comparing the pendulum equation to the mass and spring
equation, we see that the relation giving the period for a pendulum must
be:

 2π  = g
2
l
→ T = 2π
 T  l g
Also, since the equation for the mass on a spring and the equation for the
pendulum are in fact the same equation with different constants, they
must have the same solution. So the mathematical function that worked
for the mass-spring, must work for the simple pendulum too. The
distinguishing feature that makes these equations similar is that the
acceleration is proportional to the displacement, but with the opposite
sign. This is the unique feature that leads to simple harmonic motion
(SHM).

Physics 206 Fall 2002


Block 9 7

Before going any further with the analysis of SHM, it is useful to


investigate its general properties This is what we will now do.
The general solution to the equation
A more general way to write this equation that emphasizes its
applicability to all cases of SHM is as follows: We will arbitrarily use the
symbol “y”, but we just as well have used θ, or x, or any other symbol.

 2π 
2
d2
2
y(t) = -   y(t)
dt T
The general solution in standard form is,
2π t
y(t) = Asin  + φ
 T 
What is the meaning of the constants? The sine function goes
through a complete cycle every 2π radians. This occurs each time t
increases by an amount T. Thus T is the period of the oscillations, as
was mentioned before. The reciprocal of the period is the frequency of
oscillations, f:
1
f =
T
T is the time required to complete the cycle, while f is the number of
cycles per second. T is measured in seconds; f, in reciprocal seconds
(1/s), which are called hertz and abbreviated Hz.
The maximum value of the sine function is +1 and the minimum
value is -1. Thus, A is the amplitude of the oscillations. That is, the
maximum value of y is +A and the minimum value is -A .
The angle φ is determined by the value of y at the particular time
t = 0. If y has its maximum value at t = 0, then φ has to be 90˚ or π/2
radians. We say that φ depends on the “initial conditions”. The angle
φ is often referred to as the phase angle. By including the phase angle,
we can make the sine function fit any particular physical situation.
Without the phase angle, we would always have to start timing the
oscillation when the position had the value zero. By including the phase
angle, we have a perfectly general solution.
The solutions we have written down describe the position as a
function of time for any object vibrating in simple
harmonic motion. They give the specific time dependence
of the position of whatever it is that is vibrating. The three
constants depend on the particular situation.
Let’s explore our solution to SHM further. The
equilibrium value of y is the value y has when no
oscillation is occurring. For the way we have written the
solutions, this value is zero. Thus, the amplitude is the
change in y that occurs in going from the equilibrium
value to the maximum value of y.
This picture is for the phase angle φ = 0.
Changing the value of the phase angle φ shifts the curve
sideways. Compare the this plot to the plot of the sine

Physics 206 Fall 2002


Block 9 8

π
solution with φ= radians, in the graph to the right. You
4
can see that they are just the same except for a sideways
shift.
To summarize, then, the solutions we have written
down describe the position as a function of time for any
object vibrating in simple harmonic motion. They give
the specific time dependence of the position of whatever it
is that is vibrating. The three constants, A, T and φ
characterize the motion and depend on the particular
situation.
The period, T, or frequency, f, of an oscillating
system is determined by the constants appearing as the
coefficient of the linear term of the differential equation when written in
standard form:

 2π 
2
d2
2
y(t) = -   y(t)
dt T
For a mass on a spring, we found that,

 2π  = k
2
m
→ T = 2π
 T  m k
For the pendulum,

 2π  = g
2
l
→ T = 2π
 T  l g
Note that you do not need to write out the solution of the differential
equation to get the frequency or period. It comes directly from applying
Newton’s 2nd law and “reading off” the constants.
Energy for a mass hanging from a spring
As a mass hanging from a spring oscillates, very little mechanical
energy is converted to thermal energy or sound, so we expect the
mechanical energy to remain essentially constant for many periods.
The potential energy is defined in terms of the work required to
move the spring, and has the value,
PE = 12 ky 2
assuming that the origin is placed at the equilibrium value of y and that
the potential energy is defined as zero there.
As usual the kinetic energy is,
KE = 12 mv 2
Let's substitute the general solution for y(t) into these expressions:
2π t
PE = 12 ky 2 = 12 kA2 sin 2  + φ
 T 
To calculate the kinetic energy, we differentiate the expression for y(t) to
get v(t):
Physics 206 Fall 2002
Block 9 9

dy 2 π 2 πt
v(t) = = Acos + φ
dt T  T 
so that
2π 2π t
2
KE = 12 m  A2 cos2  + φ
 T   T 


2
But for a mass spring system,   = . We can substitute this
k
 T  m
relation into the equation above, finally getting the expression for the
kinetic energy:
2π t
KE = 12 kA2 cos2  + φ
 T 
We note that the maximum values of the PE and KE are the same.
Also, the time average values of the KE and the PE are the same: 1/2 of
the maximum values, since the average of sin2 or cos2 is just 1/2. Last
semester we could merely say this was plausible in our discussion of
equipartition of energy. Now we see why we were justified in saying the
average energy in the KE mode is the same as the average in the PE
mode.
When we add the kinetic energy to the potential energy to get the
total energy, we notice that we have the sum (sin2 + cos2) which always
has value unity. Hence the total energy is,
TE = 12 kA2
There are two things that are important about this equation,
(1) the total energy is a constant just as we expected, and
(2) the total energy is proportional to the square of the amplitude.
This is a characteristic of all kinds of oscillating systems and extends
even to wave motions, as will see very shortly.

Energy Graphs
The potential energy function,
PE = 12 ky 2
is a parabola when plotted as a function of y,
while the total energy, being constant, is a
horizontal line. The kinetic energy is the
difference, KE = TE - PE. But the kinetic energy
cannot be negative, since the mass is never
negative and v2 is never negative regardless of the
sign of v itself. This means that the oscillation is
limited, and can go from ymax to ymin. Of course
this just reinforces what we already know, since
ymax=A and ymin= - A.

Physics 206 Fall 2002


Block 9 10

But for any object that oscillates about an


equilibrium position, even when the force law is not as
simple as F=-ky, this graphical analysis provides an
accurate description for small oscillations. As an
example, look at the plot below:

This is the shape of the potential energy between two


bound atoms that we encountered last semester. Even
though the potential energy function is not a parabola over
all distances, near the minimum it is approximately so, and
therefore the oscillations of this system will be simple
harmonic if they are at small amplitude about the
minimum in the potential energy curve.
Thus, we can make a very strong statement:
essentially every system that vibrates, does so in SHM for
small amplitudes of vibration. All molecules, including those in our
bodies, and all atoms in solids move essentially in simple harmonic
motion.
Applying our results to the universe!
Let’s now consider our model for matter. For liquids and solids,
we picture molecules that are bound to each other as if they have little
springs attaching them together. They bounce around, and in liquids
tumble and change positions. What is the nature of the oscillations these
molecules undergo? And what happens when we exert external forces on
the matter and bend it or compress it. How then does the matter respond
on a macroscopic scale? What we saw in the discussion above on SHM,
is that if the restoring force is proportional to the displacement, but in the
opposite direction, then SHM results. This will always be the case if the
displacement from equilibrium is sufficiently small. The result is that
everything, on both the atomic scale and macroscopic scale, tends to
vibrate in SHM. This is very nice, because we know the solution! The
period of oscillation depends on factors like the mass of the particles or
object being considered and the strength of the restoring forces. If we
can identify the forces acting and write down Newton’s 2nd law, then for
small oscillations, we have the problem solved!
Think about what we have accomplished. For any kind of matter,
we know how to go about finding its vibration frequencies when
subjected to external forces as well as how its internal parts oscillate. We
have an approach that will solve zillions of problems!

Waves–What Are They?


Some Primitive Wave Concepts
A wave is a particular, although very common, type of internal
motion of a medium (material substance). In order for material waves to
exist, there must be restoring forces in the medium that cause a displaced
portion to return to its equilibrium position. In a material wave, the
particles of the medium oscillate about their equilibrium positions in a
particular kind of spatially and temporally “organized way”. The
picture of expanding rings of ripples when a rock is tossed into a pond is
a useful example to recall when we think of material waves. In this wave,
various parts of the surface of the water vibrate in a certain organized
Physics 206 Fall 2002
Block 9 11

fashion. Restoring forces exist to cause the surface of the water to


always return to the equilibrium depth.
The oscillations may be spread throughout the medium and recur
continuously at each point (repetitive wave) or the oscillations may exist
for only a limited time at each point (pulse-type wave). Our example of
expanding ripples is of this latter type.
Waves in a medium are started by outside forces that act on some
of the particles in the medium to start them oscillating. Outside forces are
not required to “keep a wave going” once it has been started. In our
example of the surface water wave, the rock dropped into the pond is the
outside disturbance that starts the wave motion. Once started, the wave
continues “on its own.” We don’t have to continue to drop rocks into
the pond to keep the ripples moving.
More Refined Definition of a Wave
A wave is the movement in specific directions of the spatial and
temporal organization of the oscillations of the particles of a medium
about their equilibrium positions.
Let’s look at each of the italicized and bold words in this
definition more closely. A wave is the movement of something. But
what is it that moves? Think about our example of the expanding ripples
in the pond. It is the ripples that move, not the water itself. The ripples
are a particular type of spatial and temporal organization of the
oscillations of the surface of the water. If we focus on a leaf floating on
the pond as the ripples pass, we see the leaf bobs up and down, i.e.,
oscillates. The water surface oscillates. But the oscillations are not the
wave. Rather it is the way the oscillations of the surface are organized in
space and in time that is the wave.
Some Properties and Characteristics
In most material waves we typically encounter, the spatial
organization of the oscillations remains fairly constant, at least over short
distances of travel. In our example, the ripples look similar as they
expand away from the initial disturbance. Over greater distances,
however, we notice changes. The ripples seem to die out as the radius of
the circle they make increases. In some waves, the spatial organization
may remain constant over very long distances, as in low frequency sound
waves at certain depths in the ocean.
Material waves provide a mechanism for transferring energy over
considerable distances, without the transport of the material medium
itself.
All waves are characterized by certain independent parameters.
Here is a brief introduction to them:
(1) Amplitude of oscillations. Amplitude is controlled by the
magnitude of the forces that start the wave.
(2) Speed of wave. Speed of a wave is determined by the properties
of the medium. Speed does not depend on the amplitude of a wave, at
least for sufficiently small amplitude waves. Neither does the speed
depend on whether the wave is a pulse or a repetitive wave. Often,
however, there is a rather small, but detectable, dependence of the
speed on the frequency of a repetitive wave (see below).
Physics 206 Fall 2002
Block 9 12

(3) Direction of particle oscillations with respect to direction of wave


travel is controlled by the direction of forces starting the wave.
Longitudinal polarization: oscillations are in the direction of wave
velocity.
Transverse polarization: oscillations are in the direction
perpendicular to wave velocity.
Repetitive waves are also characterized by one other independent variable:
Frequency of oscillations of the particles. The frequency is
controlled by the forces starting the wave.
Repetitive waves also have a dependent variable, the repeat distance or
wavelength, that characterizes the spatial organization of the oscillations.
Wavelength, a dependent variable, is related to the independent variables
frequency and wave speed:
λ = vwave / f
Here λ is the wavelength, vwave is the speed of the waves and f is their
frequency.
Example: wave on a stretched string
A stretched string is one example of a medium that has restoring
forces. Both transverse waves and longitudinal waves are possible on a
stretched string or wire. The speed, vwave, of transverse waves on a
stretched string depends on the properties of the string that affect its
elasticity and its inertial properties. For a string (rope, wire, cable) that is
thin compared to its length, the relation connecting the wave speed to the
string properties is
T
v wave =
µ
where T is the tension in the string and µ is its mass per unit length.
Note that the speed is independent of the time and, if the string is
homogeneous, then it is independent of position as well. Notice also that
the wave speed does not depend on how long the string is, or the
frequency of the wave (if it is a repetitive wave) or the shape of the pulse
(if it is a pulse-type wave).
If we “ride the wave” we are moving along with the wave; we are
moving with the wave with a velocity equal to the velocity of the wave. As
we “ride the wave”, both time and position change. But note that no part
of the medium moves with this velocity.
Wave Pulses
Waves can come in many shapes. One of the simplest is a wave
pulse. It's what you produce if you shake the end of a rope up and down
once.

Repetitive Waves
Repetitive waves can have many different shapes. One of the
simplest to deal with looks like a sine or cosine function. Such waves
Physics 206 Fall 2002
Block 9 13

are called harmonic or sinusoidal waves, and are


generated by oscillators moving in simple harmonic
motion. For example, if you hold one end of a rope and
jiggle it up and down in simple harmonic motion, you
will generate harmonic waves. If you were to take a
picture of the waves, it would look like this:

Here x is the distance along the rope, and y is


how far the rope moves sideways as the wave passes
through. The curve is the shape of the rope at the instant
the picture was taken.
The location of the rope when no wave is present
is called the equilibrium position, and is the horizontal
line in the picture.
The distance between the equilibrium position and the top of a
crest is called the amplitude and is designated here by the letter, A. It is
also the distance between the equilibrium position and the bottom of a
trough.
The distance occupied by one full oscillation is called the
wavelength and is represented by the Greek letter λ ("lambda"). It is
the distance between adjacent crests of the wave or the distance between
adjacent troughs.
If, instead of taking a picture of the wave, we paint a red dot on the
rope and then plot the position of the red dot as a function of time, we
find that it moves in simple harmonic motion:

Here T is the period of the wave motion. Just as with


simple harmonic motion, the period is the reciprocal of
the frequency:
1
T=
f
T is the time between the arrival of adjacent crests of the
wave, while f is the number of crests that pass by per
second.
This is called a one dimensional wave because is
moves only in one direction, here taken as the x-axis.
Waves on ropes, water waves in long, narrow channels
and sound waves in pipes are one-dimensional.
The mathematical wave function
We have said that a wave is the movement in specific directions of
the spatial and temporal organization of the oscillations of the medium.
For a harmonic, repetitive wave moving in one dimension, we can
represent the wave mathematically as

y( x, t ) − y0 = A sin 2 π ± 2 π + φ  = A sin Φ( x, t ) .
t x
(1-1)
 T λ 

Physics 206 Fall 2002


Block 9 14

where A is the amplitude of particle oscillation, T is the period, and λ is


the wavelength.
Note: We will number the important mathematical relationships in the
remainder of this block and continue with the same numbering scheme
in the next block, since we are talking about the exact same equations.
If we are talking about a transverse wave on a stretched string,
then y(x,t) represents the displacement of a section of the string from
equilibrium in a direction perpendicular to the direction it is stretched. If
we are talking about a sound wave in a pipe, then y(x,t) would represent
excursions of the air pressure in the pipe about its equilibrium value. The
variable x is the position along the dimension in which the wave is
traveling. The constant phase factor, φ, serves the same function here as it
did in describing simple harmonic motion. Now about the "± " sign: If
we choose the “-” sign, the motion of the wave is in the positive x-
direction; if the "+", in the negative x-direction.
Note that this expression is very similar to the mathematical
description we developed for the motion of a particle vibrating in simple
harmonic motion. The first term in the argument of the sine function,

t , is exactly the same as we had for simple harmonic motion. The
T
constant phase factor, φ, serves exactly the same purpose as for SHM,, to
give the proper value of y at t = 0 and x=0. The only thing new is the

term x . Note the similarity of this term to the term involving the
λ
time. This term involving x and λ gives the change in phase as we look
along different values of x. The phase goes through a complete cycle of
2π radians each time x increases or decreases by an amount equal to the
wavelength, λ.
The expression we’ve written above, with the arbitrary phase
constant φ, is perfectly general, and is sufficient to represent any one-
dimensional harmonic wave.
For non-harmonic waves (waves that can’t be represented with a
single sine function) and waves that don’t “go on” forever, and for two
and three dimensional waves, the mathematical representation gets a little
trickier; we won’t pursue it further here.

Important Points

The total phase of a wave


Everything inside the parentheses, the argument of the sine
function in the wave function, adds up to some angle. Let’s use the
symbol Φ for this angle. The value of the angle Φ depends, of course, on
both the time and position along the wave as well as the various constants.
2π t 2π x
Φ(x,t) = ± +φ
T λ
The wave function then takes the very simple looking form
y(x,t) = AsinΦ (x,t)

Physics 206 Fall 2002


Block 9 15

The quantity Φ is called the total phase of the wave, and is a function of
both x and t, in contrast to the phase angle φ, which is a constant.
Because the sine function is periodic with period 2π, there is no
distinction between Φ and Φ + 2π, or Φ + 2πn, where n is an integer.
Every time Φ increases by 2π, the sine function repeats. For constant x,
this occurs each time t increases by an amount equal to the period T. For
fixed t, it occurs every time x increases by an amount equal to the
wavelength λ . The period and wavelength play identical roles in the wave
function, T with respect to t, and λ with respect to x; they determine how
often the wave repeats in time and space.
The velocity of a wave
Imagine what it would mean to move along with the wave so that
the value of y(x,t) remained constant. Surfers do this as they “ride a
wave”. If we rode along the wave on a “crest”, for example, then y
would be at its maximum value, i.e., y would be equal to the amplitude A,
and Φ would have the constant value π/2. Note that even though both t
and x are changing, the phase angle Φ remains constant.
When we “ride the wave” we are moving along with the wave
with a velocity equal to the velocity of the wave. As we “ride the wave”,
both time and position change. But they change in a particular way–a
way that keeps the phase angle Φ constant (since we are constantly at the
same point on the wave, rather than moving up and down along the wave).
The wave velocity is that value of dx/dt that keeps the phase angle
constant.
We can solve for x in the expression for Φ above.
± x = -t λ/T - φλ/2π + Φ λ/2π
Then differentiating with respect to t, remembering that Φ is now a
constant, gives us the wave velocity. (Note that the square of ' ± ' is '+'
and that '- ± ' is ' m '.)
dx λ
wave velocity = =m
dt T
vwave = |dx/dt|= λ/T
The ' m ' symbol tells us that the velocity is negative if we chose the '+' in
the wave function, and positive if we chose the '-'. That means that the
wave moves in the positive x direction if we chose the “-” sign in the
wave function, and in the negative direction if we chose the “+” sign.
Note that the wave speed is independent of the fixed phase angle φ as well
as what value the total phase angle Φ happens to have. That is, it doesn’t
matter that we ride along on a crest of the wave, at the trough, or anywhere
in between. To stay at the “same spot” on the wave, we have to move
with the speed given by the simple expression λ/T.
We could have expressed the wave speed in terms of the
frequency, f, instead of the period, T. Since f=1/T,
λ
v= = λf
T . (1-2)
It is sometimes easy to get confused about which of the quantities in
Eq. (12) are dependent variables and which are the independent variables.
Physics 206 Fall 2002
Block 9 16

The properties of the medium (mass per unit length and tension for a
transverse wave on a stretched rope, for example) determine the wave
velocity, v. The object that starts the wave and gives it its energy (e.g., the
arm that shakes the end of the stretched hose) directly determines the
period. Thus, the wavelength, λ, is the dependent variable, given by
Eq. (1-2).
Graphical representation of a harmonic wave
Now one of the tricky things about the solution to the wave equation
expressed in Eq. (11) is that it is a function of both space (the distance
along the x axis) and a function of time (the value of the time variable, t).
One way to make this equation more meaningful is to think of either x or
t as being fixed at some value, so it becomes a function of only one
variable. This is exactly what we must do in order to graph the function
y(x,t) in a simple 2D graph. The following two graphs show y(x,t) as a
function of x at fixed t and as a function of t at fixed x respectively.

y(x) at a particular t

y0 + A
y0
y0 − A
λ
0 position
x

y(t) at a particular x

y0 + A
y0
y0 − A
1
T= f
0 time
t
The top graph, showing the transverse position of the rubber hose as
a function of the distance along the hose for one instant in time, actually
looks like the way the hose looks as the wave travels down the hose. That
is, the position graph is similar to a snapshot of the hose, taken at a
particular instant in time. We have this correspondence in this case
because the hose itself actually moves in a transverse fashion and the
graph is plotted in 2D with the axes perpendicular to each other. This
correspondence of the position graph to “what the medium actually looks
like” is unique to transverse waves on a one-dimensional hose, string, etc.
It certainly does not apply to 3D sound and light waves.
The lower graph shows the time dependence of a particular piece of
the hose. Each piece of the hose vibrates in simple harmonic motion
(SHM). To convey all of the information contained in Equation (11),
requires both graphs. Alternatively, the same information can be conveyed

Physics 206 Fall 2002


Block 9 17

using two position (top) graphs, each for a distinct time, or two time
(lower) graphs, each for a distinct position along the hose.
Emphasizing the displacement from equilibrium
Instead of describing the transverse position y(x,t) of a rope wave
with respect to some fixed origin, we can concentrate on the transverse
displacement ∆y(x,t) from its equilibrium position.

∆y(x,t) = Asin 2π ± 2π + φ  = Asin(Φ)


x t
 λ T  . (1-1a)
This mathematically simplifies things by concentrating on the essential
nature of waves, which is the propagation of displacements (or
fluctuations) from an equilibrium value. Graphically, this just moves the
horizontal axis to the equilibrium value of the medium. Compare the
following graph to the two preceding graphs:

transverse displacement
∆y(x) or ∆ y(t)
+A
0 position x
−A or time t

Note the change in the origin of the vertical axis of this graph
compared to the previous two graphs.

Making Eq. (1-1) less imposing


A second tricky thing about Eq. (11) or (11a) is that it looks so
complicated. But in reality, this equation is just a simple sine function.
We emphasize this by re-naming everything inside the parentheses after
the sine as the new variable Φ(x,t), and we call this the total phase of the
wave. Imagine actually making the three previous graphs. What would
you have to do? Before you could plot anything, you would have to
substitute in values for x, t, λ, T, and φ in order to calculate a total angle (in
radians). Then you would take the sin of the angle and multiply by the
amplitude. This gives you one value of y to plot one point on the graph.
What determines the value of y at a particular point? The total phase–the
thing you take the sine of, Φ (x,t). And what determines Φ (x,t)? The
values of the fixed parameters λ, T, and φ and the particular values of the
independent variables, x and t.
The phase determines everything
When we start adding waves together, we will find that what’s
important is the total phase of each of the waves, Φ1(x,t) and Φ2(x,t). You
should try to think of these as simply angles (which is what they are),
because you know a lot about how the sine of an angle behaves. In
particular, you know that you can always add 2π or 4π or -8π or any
integer number of 2πs to an angle (or subtract them), and when you take
the sine of it, you get back exactly the same thing. We say the sine
function is periodic with period 2π radians. You also know that if you
add π, or 3π or -5π to an angle and take the sine of it, you get back the
same numerical value, but with opposite algebraic sign.
Physics 206 Fall 2002
Block 9 18

Now, when we get to the application of these ideas, the interesting


thing is that there are so many ways to change the total phase of the two
waves. (This is also what can make it seem like we are dealing with a
large number of unrelated phenomena.) There are the three wave
parameters for each wave (λ, T, and φ) and the two independent sets of
variables, x and t, (or origins for measuring x and t, if we prefer to think of
x and t themselves as common variables) for a total of ten things that
could change. All of the various phenomena we are about to discuss can
be understood as particular cases where we change this or that parameter
or variable. But in the end, we still end up with a specific total phase
angle for each wave. The almost limitless number of possibilities comes
from the many ways the total phase of each wave can be changed. The
simplicity comes from the fact that in the end, it is still just a change in the
total phase angle that determines what happens.
The full meaning of the preceding paragraph will come to you as you
perform the DL activities and work on the FNTs. Remember to try and
think of Eq (11) as a simple sine function of an angle, the total phase
angle, Φ, and that the interesting phenomena occur as a result of things
that change the total phase angle.
The velocity of the particles
What is the motion of an individual particle of the medium? In
particular what is the velocity of a particle? Let’s imagine a transverse
wave on a string. The wave is traveling in the x-direction and the
vibrations are in the y- direction. Then the general wave function as we
have written it applies. What is the motion of a small length of the
stretched string? Well, the wave function tells us the position of the
string about its equilibrium position. To get the velocity of a small length
of the string located at some distance x from the origin, we can take the
time derivative of y(x,t),
∂y 2π 2π t 2π x
vparticle= =A cos  ± + φ
∂t T  T λ 
where we keep x constant at our position when taking the derivative
(technically this is called a partial derivative). Any given piece of the
string is oscillating with SHM. The phase of oscillation depends on
where along the string we look, i.e., on x. But each little piece of the
string moves in SHM with the same amplitude as every other piece.
Keep in mind that vparticle is not vwave. In a string it is the up and
down velocity of a point on the string, not the sideways, or transverse
motion of the wave.
The energy of the wave
Since we have an expression for the velocity of the particles, we
can immediately write down an expression for the energy contained in
our one dimensional wave. Let’s let µ be the mass per unit length of the
string. Then the kinetic energy per unit length of string is
KE/unit length = 1
2 µ vparticle2
There will also be a potential energy associated with each unit length of
string. As the piece of string oscillates, energy is continually transferred
back and forth between potential and kinetic. However, when the piece of
string is at its equilibrium position, all of the energy is kinetic, so the total
energy per unit length of string is
Physics 206 Fall 2002
Block 9 19

Tot E/unit length = 1


2 µ vparticle2 (at y = 0)
We want the velocity when y equals zero. This is just the maximum
velocity, which occurs when the total phase angle is zero, so that the
cosine has a value of one. Thus

vparticle max = A and
T

2
Tot E/unit length = 1
µA  
2
2
 T 
= 12 µ A2 (2 π f )2
This is the energy per unit length of the wave. Note that it depends on the
square of the amplitude and frequency. This is true for all one-
dimensional waves. We will discuss transfers of wave energy to other
systems when we get to sound waves.
Applying our wave model to sound and light
Another point that we want to discuss before getting too involved
with adding waves is to what extent our wave model applies to real
physical phenomena. We previously talked about Eq. (1-1) representing
a continuous harmonic wave on a stretched rubber hose. As a beginning
scientist, you should be getting into the habit of automatically asking
yourself questions such as, “How good is this model for a rubber
hose?” “Under what conditions does the model ‘break down’?” Our
experience with vibrating objects helps us here. In order to have
vibrations, there must be a restoring force. We previously saw that for
physical systems involving restoring forces that increase with
displacement from equilibrium, the SHM model became a good fit even
to forces that are nonlinear in displacement as the amplitudes of vibration
are made sufficiently small. That is, for small oscillations, the SHM
model fits essentially all real physical phenomena involving oscillation
about an equilibrium position. Since a wave involves the coupled
oscillations of many separate oscillators, we might expect to find the same
type of applicability. This is indeed the case. If the amplitude of the
disturbance in the media about its equilibrium value is sufficiently small,
the harmonic wave model is an excellent model. That is, predictions made
using the model do indeed correspond to physical measurements in real
media.
So, from the discussion of the preceding paragraph, we would expect
the wave solution, expressed in Eq. (1-1), to accurately describe the
excursions of a rubber hose as a harmonic wave travels down the hose, at
least as long as the amplitude of the wave is not too great. But what about
other kinds of phenomena we would like to apply our wave model to? In
particular, what about sound and visible light, and microwaves, and radio
and TV waves, and short wavelength cellular telephone waves? What is
involved in applying the wave solution expressed in Eq. (1-1) to these
phenomena? In the first place, what does the symbol y now represent?
There is no longer a stretched rubber hose that vibrates. What does
vibrate?
A wave model of sound
Let’s first begin with sound. If we are dealing with vibrations of the
air in long, narrow tubes, trombones, flutes, or trumpets, for example, we
Physics 206 Fall 2002
Block 9 20

basically have 1D phenomena, and our 1D wave function, y(x,t) described


by Eq. (1-1) is a good model for how the air vibrates when traveling
harmonic waves are present. But vibrating air is very different from the
vibrations of a stretched rubber hose. We need a good mental picture of
exactly what vibrates and how it vibrates. We know something about air.
The particle model you learned about in your chemistry course and which
we extended in Physics 205 tells us that the air molecules are in random
motion, moving rapidly about, but that in equilibrium, there will be well
defined average values of quantities such as particle density, pressure, and
temperature. Sound in air involves the oscillation of the average value of
particle density (and resulting pressure) over distance scales much larger
than the mean distance between particles. So what vibrates in a sound
wave in air? The average value of particle density (and pressure). Thus,
the y in Eq. (11) represents the average particle density (or pressure) of
the air inside a tube as we move along the length of the tube in the x
direction.
Let’s rewrite Eq. (11) explicitly using the symbol P(x,t) to represent
the absolute pressure of the air inside a long tube.

P(x,t) − Patm = Asin  2π ± 2π + φ  = Asin(Φ),


t x
(1-3)
 T λ 
where P(x,t) is the absolute pressure of the air at a given position x
along the tube, and at a time t. P atm is the equilibrium pressure (i.e.,
atmospheric pressure), and A is the amplitude of the pressure fluctuation
(gauge pressure) from equilibrium. Using Eq. (1-3), we can graphically
represent the air pressure P(x) of a sound wave in a tube for a range of x
positions at a specific moment in time.

P(x) at a particular t

P atm + A
P atm
P atm − A
λ
0 position
x
This definitely does not resemble a picture of a sound wave (cf. QP 1-4);
it is only a graphical representation of the pressure fluctuations in
harmonic waves in the air in a long tube.
We can also graphically represent the air pressure P(t) of a sound
wave in a tube for a range of times at a specific point in this tube, as
shown below.

Physics 206 Fall 2002


Block 9 21

P(t) at a particular x

P atm + A
P atm
P atm − A
1
T= f
0 time
t
If you compare Eqs. (1-1) and (1-3) and the graphs associated with
each equation, you see that the mathematical and graphical harmonic
model of sound waves in the air in a long tube is exactly the same as for
waves on a stretched rubber hose. This is because both transverse waves
on rubber hoses and sound waves in air constrained to long tubes are
waves that are free to propagate along only one dimension. But there is
one very important difference. The graph of the displacement of the hose
as a function of distance along x actually looks like a snapshot of the
wave on the hose. The graph above showing the pressure fluctuations of
the air in the tube as a function of the distance x along the tube does not
“look like” anything having to do with the air molecules. It is not even a
graph of the average positions of the air molecules. It is simply a graph
of the pressure fluctuations along the tube. For most of us, this gets kind
of hard to “picture.” What are the air molecules actually doing? QP 1-4
helps you to get a handle on the relation between what the air molecules
are doing and the variation of pressure associated with a wave. But
without minimizing the usefulness of trying to get this picture, we can
understand much about sound as a wave by focusing only on Eq. (1-3)
and its corresponding graphs. Exactly what it is that vibrates is a very
interesting question, but difficult to visualize for sound. (It gets much
worse for visible light and other electromagnetic waves!) But what saves
us is that we can rely on our model to get a mental picture of the wave
phenomena, instead of relying on an actual picture of whatever it is that is
actually vibrating! What we are saying is that we can acquire a level of
comfort and security with Eq. (1-1) or Eq. (1-3), i.e., the model, that lets
us understand complicated phenomena involving sound or light without
the necessity of having a concrete picture of vibrating objects. This is an
example of the power of mathematical models for gaining an
understanding of complicated phenomena that are difficult or impossible
to envision concretely.
A wave model of light
Visible light is a fascinating phenomenon we could spend an entire
semester on (Physics 207?) in trying to fully comprehend its nature, and
modeling its multifaceted behavior, but alas, we will only get to implicitly
appreciate some of the more basic (but important) aspects of light. So
here we'll just mention a few things about light in general, and point out
some properties it has in parallel with sound and other wave phenomena.
Many phenomena involving visible light make sense if we model
light as a wave that travels at a constant speed of c = 2.998 x 108 m/s
(usually rounded to 3 x 108 m/s) in vacuum. It is a transverse excitation,
meaning that whatever it is that vibrates, vibrates like a stretched rubber
hose, perpendicular to the direction of wave propagation. For these
Physics 206 Fall 2002
Block 9 22

phenomena, light is well described by the mathematical models and


principles we have discussed so far for sound waves and waves in
general. Like all waves, light waves have frequencies determined by the
source of the waves. The human eye is biochemically engineered to
respond to light frequencies in the range of 4.3 × 1014 Hz to 7.5 × 1014
Hz (as with all subjective experiences, this range will be slightly different
for different people). Within this range, light is subjectively experienced
as color—red corresponds to the lower frequency limit; violet
corresponds to the upper frequency limit; and all other colors correspond
to frequencies within this visible light range. Light frequencies just lower
than 4.3 × 1014 Hz are called infrared (i.e., “lower than red”), and even
lower frequencies correspond to microwave, then TV, FM, and AM radio
signals. Light frequencies just higher than 7.5 × 1014 Hz are called
ultraviolet (“beyond violet”), and even higher frequencies correspond to
x-rays and gamma rays. Taken together, all of these waves are referred to
as electromagnetic waves or electromagnetic radiation. Visible light
occupies one small part of the frequency range. As with sound waves
and our ears, there are quite a lot of different types of light frequencies
that cannot be perceived by our eyes. Just because we cannot detect and
subjectively perceive these different frequencies of light as color does not
mean that they do not exist–we only need to construct different types of
“detectors” (like the microwave detectors used in DL) that are sensitive
to these other light frequencies in order to “see” them.
Now you should be asking yourself the question, “What waves or
vibrates in a light wave?” This question puzzled physicists for many
years. If light really acts like a wave, i.e., if it can be modeled as a wave,
there must be a medium that vibrates. Physicists called this supposed
material medium the ether. However, the ether could never be directly
detected and it was eventually realized that light required no material
medium to propagate through. Light can indeed propagate through air,
water, and other “transparent mediums”, but it can also propagate
through empty space, a perfect vacuum. So if visible light (and all other
forms of electromagnetic radiation) don’t involve the vibrations of a
material medium, what does vibrate or wave? The answer is that it is a
combination of electric and magnetic fields that oscillate. Later in this
course we will try to develop a useful mental picture of oscillating electric
and magnetic fields, but for now, we will just except that there is
something that vibrates. Then, when we have a need for a one
dimensional light wave, we can use Eq. (11) to describe the oscillations of
the electric and magnetic fields.
Other ways to model light
It turns out that we cannot exclusively describe light as just a wave
because light is not just a wave. You have seen the quantum model for
light, where light energy is transferred by photons, discrete packets of
energy. Whether we model light as a wave or a photon will usually
depend on whether we are concerned with macroscopic or quantum-scale
behavior of light. (Just for your information, sound can also be described
as phonons, discrete packets of energy on the atomic level.)
As Physics 206 progresses you will be discussing these various
models for light, most of which are complementary in nature, yet some
have properties that will seem contradictory. There will be no easy
resolutions to the problem of adequately describing light, but that
shouldn't mean light is something that cannot be well understood; it just
Physics 206 Fall 2002
Block 9 23

means we have to be careful to choose the most appropriate model for the
particular phenomenon we are dealing with.

Summary of Block 9
Simple Harmonic Motion
Mass on a spring
The differential equation for the motion is,

 2π 
2
d2
2
x(t) = -   x(t)
dt T
provided

 2π  = k
2
m
→ T = 2π
 T  m k
Here x is the distance from the equilibrium position of the mass as it
hangs motionless on the spring. The solution is,
2π t
x(t) = Asin  + φ
 T 
Simple Pendulum
The differential equation for the motion is,

d 2θ  2π  2
=- θ (t)
dt 2  T 
where

 2π  = g
2
l
→ T = 2π
 T  l g
Here θ is the angle the pendulum makes with the vertical. The solution is,
2π t
θ (t) = Asin  + φ
 T 
Frequency
The reciprocal of the period is the frequency of oscillations, f:
1
f =
T
Energy
The kinetic energy of the motion of a mass hanging from a spring is,
2π t
KE = 12 mv2 = 12 kA2 cos2  + φ
 T 
The potential energy associated with the restoring force of the spring is,
2π t
PE = 12 ky 2 = 12 kA2 sin 2  + φ
 T 
Physics 206 Fall 2002
Block 9 24

The sum of these two is,


TE = 12 kA2
Note that the total energy is constant.

Waves
The Wave Function
The wave function of a disturbance, y, is,
2π t 2π x
y(x,t) = A sin  ± + φ
 T λ 
where A is the amplitude of particle oscillation, T is the period, λ is the
wavelength, and φ is the phase factor.
The overall phase of a wave, Φ, is defined as everything inside the
parentheses:
2π t 2π x
Φ (x,t) = ± +φ
T λ
The wave function then takes the very simple looking form
y(x,t) = AsinΦ (x,t)
Wave Velocity
The wave velocity, vwave, is
vwave = |dx/dt|= λ/T
In terms of λ and T, the wave velocity is,
vwave=λf=λ/T
Particle Velocity
The velocity of the particles in the medium is the particle velocity:
∂y 2π 2π t 2π x
vparticle= =A cos  ± + φ
∂t T  T λ 
Energy in the wave
Waves carry energy. For a wave in a string,


2
Tot E/unit length = µ A  
1 2
2
 T 
Here µ is the mass per unit length of the string.

Physics 206 Fall 2002

You might also like