Sunteți pe pagina 1din 50

Simple Harmonic Motion

Simple harmonic motion is typified by the motion of a mass on a spring when it is subject to
the linear elastic restoring force given by Hooke's Law. The motion is sinusoidal in time and
demonstrates a single resonant frequency.

Index

Periodic
motion
concepts

Motion equations Motion calculation Frequency calculation Motion sequence visualization


Damped oscillation Driven oscillation

Go Back
HyperPhysics***** Mechanics R Nave
Simple Harmonic Motion Equations
The motion equation for simple harmonic motion contains a complete description of the
motion, and other parameters of the motion can be calculated from it.

The velocity and acceleration are given by Index

Periodic
motion
concepts

The total energy for an undamped oscillator is the sum of its kinetic
energy and potential energy, which is constant at

Energy transformation in periodic motion

Go Back
HyperPhysics***** Mechanics R Nave
Simple Harmonic Motion Calculation
The motion equations for simple harmonic motion provide for calculating any parameter of
the motion if the others are known.

Index

Periodic
If the period is T = s motion
concepts

then the frequency is f = Hz and the angular frequency = rad/s.

The motion is described by

Displacement = Amplitude x sin (angular frequency x time)


y = A x sin ( x t )
m m rad/s s
= x sin ( x )

Any of the parameters in the motion equation can be calculated by clicking on the active word
in the motion relationship above. Default values will be entered for any missing data, but
those values may be changed and the calculation repeated. The angular frequency calculation
assumes that the motion is in its first period and therefore calculates the smallest value of
angular frequency which will match the other parameters. The time calculation calculates the
first time the motion reaches the specified displacement, i.e., the time during the first period.
Go Back
HyperPhysics***** Mechanics R Nave

http://hyperphysics.phy-astr.gsu.edu/hbase/shm.html
Reference:

What Is Simple Harmonic Motion?

By Jim Lucas, Live Science Contributor | October 29, 2015 02:11am ET

The vibration of a guitar string is an example of simple harmonic motion.


Credit: Joshua David Treisner | Shutterstock

When a musician strums a guitar, the vibration of the strings creates sound
waves that human ears hear as music. When a guitar string is plucked, it
moves a certain distance, depending on how hard the guitar player strums.
The string returns to its starting point and travels nearly the same distance
in the opposite direction. The vibrational energy of the string is dissipated in
the form of sound. This causes the distance the string moves, or the
amplitude of the vibrations, to decrease gradually. The volume of the sound
fades until the string eventually falls silent.

The guitar string is an example of simple harmonic motion, or SHM. SHM can
be seen throughout nature. It describes the vibration of atoms, the variability
of giant stars, and countless other systems from musical instruments to
swaying skyscrapers.

Princeton University's WordNet defines simple harmonic motion as periodic


motion in which the restoring force is proportional to the displacement.
Periodic means that the motion repeats at a constant, predictable rate. The
part about the restoring force being proportional to the displacement simply
means the farther you push something, the harder it pushes back.

SHM also describes the motion of a ball hanging from a spring. The ball is
subject to a linear elastic restoring force, according to Georgia State
University's HyperPhysics website. Drawn out on a graph, the up-and-down
motion of the ball over time can be illustrated as a smooth, repetitive
oscillation, or sinusoidal wave. The shape of the wave can be calculated
using Hooke's Law.

Robert Hooke was a British scientist, mathematician and architect who was
interested in many aspects of science and technology, from astronomy to
microbiology. Studying objects under a magnifying glass, he was the first to
use the word "cell" to describe the components of plant tissues, according to
the Physics Hypertextbook. He also studied clocks and timekeeping, and in
1675 developed a theory of elasticity or springiness as a way to regulate
pendulum clocks. In short, the theory says, "Extension is directly proportional
to force."

Mathematically, Hookes Law is expressed as F = kx, where F is the force, x


is the change in length from the springs relaxed or original length, and k is
the characteristic spring constant that specifies the amount of force required
to stretch or compress the spring a certain distance. The minus sign
indicates that the restoring force is in the opposite direction of the
displacement.
A ball on a spring is the standard example of periodic motion. If the displacement of
the mass is plotted as a function of time, it will trace out a sinusoidal wave.
Credit: Georgia State University

Circular motion

There is a close connection between circular motion and simple harmonic


motion, according to Boston University. Consider a point on the rim of a disk
as it rotates counterclockwise at a constant rate around a horizontal axis. If
we plot only the vertical position of the point as the disk turns, it produces a
sinusoidal graph. This is exactly the same graph as we get if we plot the
position of a mass on a spring bouncing up and down in simple harmonic
motion as a function of time.

Pendulums

Simple pendulums behave much like harmonic oscillators such as springs.


However, the period of a pendulum is determined not by its mass but by its
length. Another significant difference is that in the case of a pendulum, the
restoring force is provided not by a spring but by gravity. Since gravity is
pulling the pendulum vertically downward and not back along the arc in the
opposite direction of its motion, the restoring force is a somewhat
complex trigonometric function. Nevertheless, pendulums closely
approximate simple harmonic motion provided they dont swing more than a
few degrees from their resting point.

Damped harmonic motion

All mechanical systems are subject to damping forces, which cause the
amplitude of the motion to decrease over time. These forces can include
frictional forces between moving parts, air resistance or internal forces such
as those in springs that tend to dissipate energy as heat.

In simple harmonic motion, the damping force is generally proportional to


the velocity of the oscillating mass, according to the State University of New
York Stony Brook. This force will eventually bring an oscillating system to a
halt, and if it is great enough, it can actually stop oscillations before they can
start.

A weight on a spring bouncing in air will continue bouncing for quite a long
time, but not forever. Air resistance and internal resistance in the metal
spring will eventually dissipate the kinetic energy of the system and bring it
to a stop. However, if the weight were suspended in a pot of cold molasses,
and the spring is stretched and released, the weight might only return slowly
to its rest position without bouncing above it even once. We consider the
weight bouncing in air to be underdamped, and the weight moving slowly
through molasses and never bouncing even once to be overdamped. If the
system is exactly on the cusp between these two conditions, it is said to be
critically damped.

We know that in reality, a spring won't oscillate forever. Frictional forces will
diminish the amplitude of oscillation until eventually the system is at rest,
according to the University of California Santa Cruz. The amplitude of an
underdamped oscillator undergoes an exponential decay, meaning that after
a certain time, the amplitude of the oscillations will decrease by half, and
after that same time period, it will decrease again by half. One familiar
example of this is the decaying sound of a bell.

In many cases, damping is desired to prevent uncontrolled bouncing. Perhaps


the most familiar example of this is an automobile suspension system. The
wheels are mounted on springs so they can move up and down in response
to bumps and dips in the road while the body of the car remains relatively
level. However, without damping, the car would begin to bounce up and
down uncontrollably. To prevent this, cars have shock absorbers to damp the
movement of the springs by adding a frictional force to the suspension
system.

Driven harmonic motion

When it is desirable for harmonic motion to persist, damping is a problem


that must be overcome with a driving force. Consider the pendulum in a
grandfather clock. At its farthest point in each direction, an escapement
mechanism, powered by the gravitational energy of a slowly descending
weight, gives the pendulum a tiny nudge that is sufficient to overcome the
damping effect of mechanical friction and air resistance and maintain the
pendulums motion.

One manifestation of driven harmonic motion is coupled resonance, or


sympathetic vibration. This is a vibration that is produced in one body by the
vibrations of exactly the same period in a neighboring body. One familiar
example of this is a playground swing set. If it has two swings of the same
length, swinging on one swing can cause the other one to start moving. The
reason for this is that the first swing causes the supporting cross bar to move
forward and back slightly with each cycle. Although this motion is usually
small, because swings are the same length, they will have the same resonant
frequency, so the driving force provided by the tiny motions of the cross bar
become amplified, causing the second swing to move noticeably. Another
example of a harmonic oscillator being driven at its resonant frequency is
how an electric guitar can sustain a note indefinitely by allowing feedback of
the amplified sound to drive the continued vibration of the string.

Sometimes, though, harmonic motion under a driving force can be harmful or


even disastrous. If this force is applied at the natural or resonant frequency
of the oscillating system at a rate that is greater than damping forces can
dissipate the energy, relatively small driving forces can build up to large-
amplitude oscillations, just because energy is continually being injected into
the system at just the right frequency, according to Boston University. This
means that if an underdamped system is driven by external forces at its
resonant frequency, the amplitude of the motion can increase until the
system fails. One of the most dramatic examples of this is the 1940 Tacoma
Narrows Bridge collapse. Strong sustained winds drove oscillations of the
bridge deck that increased in amplitude until it broke apart. Another example
of harmonic motion being driven to the point of failure is how singer can
shatter a wine glass by loudly singing a note at its resonant frequency.
Harmonic motion in real life is rarely simple. When we look at real
macroscopic oscillating systems, there can be any number confounding
variables, such as fluctuations in temperature and air pressure or wear and
tear on mechanical parts, which can affect the amplitude and period of the
motion. However, by assuming that mechanical systems are in simple
harmonic motion, we can often make suitably accurate approximations of
how these systems behave.

http://www.livescience.com/52628-simple-harmonic-motion.html

Simple harmonic motion


11-17-99

Sections 10.1 - 10.4

The connection between uniform circular motion and SHM


It might seem like we've started a topic that is completely unrelated to what we've
done previously; however, there is a close connection between circular motion and
simple harmonic motion. Consider an object experiencing uniform circular motion,
such as a mass sitting on the edge of a rotating turntable. This is two-dimensional
motion, and the x and y position of the object at any time can be found by applying
the equations:

The motion is uniform circular motion, meaning that the angular velocity is constant,
and the angular displacement is related to the angular velocity by the equation:
Plugging this in to the x and y positions makes it clear that these are the equations
giving the coordinates of the object at any point in time, assuming the object was at
the position x = r on the x-axis at time = 0:

How does this relate to simple harmonic motion? An object experiencing simple
harmonic motion is traveling in one dimension, and its one-dimensional motion is
given by an equation of the form

The amplitude is simply the maximum displacement of the object from the
equilibrium position.

So, in other words, the same equation applies to the position of an object experiencing
simple harmonic motion and one dimension of the position of an object experiencing
uniform circular motion. Note that the in the SHM displacement equation is known
as the angular frequency. It is related to the frequency (f) of the motion, and inversely
related to the period (T):

The frequency is how many oscillations there are per second, having units of hertz
(Hz); the period is how long it takes to make one oscillation.

Velocity in SHM
In simple harmonic motion, the velocity constantly changes, oscillating just as the
displacement does. When the displacement is maximum, however, the velocity is
zero; when the displacement is zero, the velocity is maximum. It turns out that the
velocity is given by:

Acceleration in SHM
The acceleration also oscillates in simple harmonic motion. If you consider a mass on
a spring, when the displacement is zero the acceleration is also zero, because the
spring applies no force. When the displacement is maximum, the acceleration is
maximum, because the spring applies maximum force; the force applied by the spring
is in the opposite direction as the displacement. The acceleration is given by:
Note that the equation for acceleration is similar to the equation for displacement. The
acceleration can in fact be written as:

All of the equations above, for displacement, velocity, and acceleration as a function
of time, apply to any system undergoing simple harmonic motion. What distinguishes
one system from another is what determines the frequency of the motion. We'll look at
that for two systems, a mass on a spring, and a pendulum.

The frequency of the motion for a mass on a spring


For SHM, the oscillation frequency depends on the restoring force. For a mass on a
spring, where the restoring force is F = -kx, this gives:

This is the net force acting, so it equals ma:

This gives a relationship between the angular velocity, the spring constant, and the
mass:

The simple pendulum


A simple pendulum is a pendulum with all the mass the same distance from the
support point, like a ball on the end of a string. Gravity provides the restoring force (a
component of the weight of the pendulum).

Summing torques, the restoring torque being the only one, gives:

For small angular displacements :

So, the torque equation becomes:

Whenever the acceleration is proportional to, and in the opposite direction as, the
displacement, the motion is simple harmonic.

For a simple pendulum, with all the mass the same distance from the suspension point,
the moment of inertia is:

The equation relating the angular acceleration to the angular displacement for a simple
pendulum thus becomes:

This gives the angular frequency of the simple harmonic motion of the simple
pendulum, because:

Note that the frequency is independent of the mass of the pendulum.

http://physics.bu.edu/~duffy/py105/SHM.html

Simple Harmonic Motion


Simple harmonic motion refers to the periodic sinusoidal oscillation of an object or quantity. Simple harmonic motion
is executed by any quantity obeying the differential equation

(1
)
where denotes the second derivative of with respect to , and is the angular frequency of oscillation.
This ordinary differential equation has an irregular singularity at . The general solution is

(2)

(3)

where the two constants and (or and ) are determined from the initial conditions.

Many physical systems undergoing small displacements, including any objects obeying Hooke's law, exhibit simple
harmonic motion. This equation arises, for example, in the analysis of the flow of current in an electronic CL circuit
(which contains a capacitor and an inductor). If a damping force such as Friction is present, an additional term
must be added to the differential equation and motion dies out over time.

SEE ALSO:Damped Simple Harmonic Motion, Harmonic Addition Theorem, Simple Harmonic--Motion Quadratic
Perturbation, Uniform Circular Motion

CITE THIS AS:

Weisstein, Eric W. "Simple Harmonic Motion." From MathWorld--A Wolfram Web


Resource. http://mathworld.wolfram.com/SimpleHarmonicMotion.html

Simple Harmonic Motion


Page 1 of 4
This section will cover the following topics:

Oscillations

Displacement

Velocity and Acceleration

Equilibrium and Restoring Forces

Inertia

Sine Waves

Damping

Forced Oscillations, Resonance and Breaking Glasses


Damping revisited

Oscillations
You may have heard the one about the opera singer breaking a wineglass with their voice...(in other versions
it's a greenhouse(?!), or even the specs worn by members of the audience). Well, if the singer could pull it off,
they would be exploiting oscillations. Later in this section we'll show you how to break a wineglass with a
loudspeaker...but first we'll try to work out some rules for things which vibrate, or oscillate. There are many
forms of oscillation in the real world - oscillations determine the sound of a musical instrument, the colour of a
rainbow, the ticking of a clock and even the temperature of a cup of tea. A mechanical oscillation is
a repeating movement - an electrical oscillation is a repeating change in voltage and current.

Mechanical oscillations are probably easiest to think about, so we'll start there. Let's think about the movement
of an object back and forth over a fixed range of positions, such as the movement of a swing in a playground,
or the bouncing up and down of a weight on the end of a spring (this last one is one of those examples which
teachers / scientists / engineers love, but which doesn't seem to have anything to do with the real world. As
we'll see, it does!!).

The pendulum and the mass-spring system are both oscillating. How can we describe their oscillation?

Displacement
The animation shows how displacement describes the distance (and also direction...displacement can be
negative as well as positive) of the object from its equilibrium position.

The displacement amplitude tells us how 'big' the oscillations are - we can use the peak value (the maximum
positive displacement from the equilibrium position) or the peak-to-peak value (the distance between negative-
maximum to positive-maximum).

The time period of the oscillation is the time taken for the object to travel through one complete cycle. We can
measure from the equilibrium position, as in the animation, or from any other point on the cycle, as long as we
measure the time taken to return to the same point with the same direction of travel.

Can you see why the first time the mass returns to the equilibrium position, we've only reached half a cycle?

Displacement on its own will sometimes do...but often it is also useful to understand the oscillation in terms
of velocity or acceleration. These three things are very similar...read on.

Velocity and Acceleration


When our oscillating object reaches maximum displacement (when it is as far from equilibrium as it can get) it
changes direction. This must mean that there is an instant in time when it is not moving...and so at the point of
maximum displacement, its speed is zero. Then it speeds up in the opposite direction, and travels fast through
the equilibrium position before starting to slow again in preparation for the next change in direction.

Velocity is just another word for speed, with the extra feature that it has direction and therefore can be negative
or positive. (If you walk backwards at 4m.p.h. - difficult without falling over - then your velocity = -4m.p.h.)

So - when velocity is maximum, displacement is zero, and when displacement is maximum, velocity is zero.
The period of oscillation is the same, but the plots for displacement and velocity do not 'line-up' in time...one is
shifted along compared to the other. This shift along in time is called a phase angle, which is an idea we'll come
back to later.

At maximum displacement the object has to come to rest, and then start moving in the opposite direction.
Imagine braking in a car (moving forwards) before immediately moving off fast in reverse - you'd feel
rapid deceleration during braking, followed acceleration in reverse gear. Like displacement and velocity,
acceleration can be negative or positive depending on direction - in this example, braking would cause negative
acceleration, as would moving off in reverse. It sounds wierd, but the point in time where the acceleration would
be (negative) maximum is the moment when the car stops moving - just as it changes direction.

Thinking about our animated mass-spring system, we can see this effect. Conversely as the object moves
through its equilibrium position, the velocity is only changing very slowly and the acceleration actually becomes
zero. This is shown below:

Note again - the period is the same, but things don't line up in time...there's some phase in there too.

All these masses and springs are OK - but you can experience this by playing on a swing. Look at these
videos, and try to work out when displacement, velocity and acceleration are 1) a maximum, and 2) zero, for
someone swinging...

Equilibrium and Restoring Forces


In our animations, the mass on a spring bounces up and down with a regular 'pattern' or waveform. Why does
this happen?

Let's think some more about playground swings. What's the first thing we do to get the swing moving?

The swing is first pulled back, and then pushed forward. It then oscillates back and forth 'on its own' until it
slowly comes to rest at its equilibrium position - the middle position between the two extremes of displacement.
This is usually the place where an object will naturally rest if no external forces are applied to it.

Once the swing has been pulled away from its equilibrium position, the force of gravity will act to bring it back.
This force will always act in a direction towards the equilibrium position, and is known as a restoring force. If the
swing is pulled higher into the air (larger displacement amplitude) then there will be a larger restoring force
acting on it, and when the pusher lets go it will travel further. This shows that the restoring force is proportional
to the distance of the swing from the equilibrium position.

The animation below shows these forces in action:

So why doesn't the swing just return directly to the equilibrium position, and stay there?

The restoring forces are large enough to make the swing over-shoot and travel 'up the other side'; this is when
a restoring force in the opposite direction takes over and pushes the swing back down. Again, the force is large
enough for the swing to over-shoot equilibrium and travel 'up the other side'. The process keeps repeating and
an oscillation is formed. If no other forces are applied to the system, the swing will keep swinging back and
forth forever, with a constant period and constant amplitude.

Does this happen in real life? (Silly question - of course not). Then why not?

In all real-life oscillating systems, damping comes into play. Damping is all about the loss of energy, often due
to resistance or friction. The air displaced by the swinger adds some resistance, and even if the swing were in a
vacuum (and the swinger were, therefore, dead) the pivots at the end of the chain generate some friction with
their supports and slowly, almost immeasurably, heat up. This uses enery, which has to come from
somewhere...and so the amplitude of oscillation decreases.

If you have no pusher, and have to propel yourself on the swing, you have to put some effort into it to keep the
swing moving. This effort is overcoming damping. If there were no damping (and also if you could switch off
your heart, brain and other energy-consumptive body functions) you could swing forever without needing to eat!

http://www.acoustics.salford.ac.uk/feschools/waves/shm.php

Inertia
Let's go back to thinking about masses and springs. Imagine a mass held between two springs as shown in the
animation below. If the mass is moved away from the equilibrium position the restoring forces provided by the
springs will make the mass oscillate back and forth in a similar manner to the playground swing example shown
on the previous page.

What happens if we change the size of the mass?

An increase in mass increases the inertia (reluctance to change velocity...i.e. reluctance to accelerate /
decelerate) of the object. (In this example we have switched off gravity, so the heavy mass does not 'sag' on
the springs. Note - massive objects have inertia even in outer space where they have no weight!). An increased
inertia means that the springs will not be able to make the mass change direction as quickly. This increases the
time period of the oscillation. The greater the inertia of an oscillating object the greater the time period; this
lowers the frequency of its oscillations.

The oscillating objects we've looked at up to now all vibrate with a rather special 'shape' or waveform. Let's
think some more about that...

Sine Waves
The animation below shows how an oscillating pattern (waveform) can be modelled by placing a marker on the
edge of a rotating disc. If we just think about the VERTICAL location of this marker (i.e. view the disc edge-on)
then we can see that it moves up and down with simple harmonic motion. It generates the sine-wave
(sinusoidal) waveform we have been using for mass-spring systems and pendulums. Objects which oscillate
sinusoidally (which = most things in practice!) are described as oscillating with simple harmonic motion.

(This 'disk' idea is very useful for thinking about phase of oscillation - we'll come back to this later).

Simple Harmonic Motion and Resonant Frequencies

Here's the maths which underlies what we've been talking about.

An oscillation follows simple harmonic motion if it fulfils the following two rules:

1. Acceleration is always in the opposite direction to the displacement from the equilibrium position

2. Acceleration is proportional to the displacement from the equilibrium position

The acceleration and displacement are linked by the following equation:

Acceleration = - 2 * x
Here is called the angular frequency of oscillation, and is given by 2 / T or 2 f

T is the period of oscillation (s), f=1/T = frequency of oscillation (Hz) and x is the displacement (m).

Using Newton's 2nd Law (F=ma) we can show that the Force on the object due to inertia will be:

F= -m 2 x

Using Hooke's Law for springs(F=-kx), we know that the force on the oscillating mass due to the springs is
simply

F=-kx

These two forces are always in balance, so

m 2 x - kx=0

From this we can find the resonant frequency:

2=(kx) / (mx) = k / m,

= sq.root (k / m)

f = 1/ 2 * { sq.root (k / m) }

This might look a little 'tricky'...but it is a massively useful equation!

Damping
When we were talking about playground swings, we mentioned damping - a loss of energy from, in that
example, movement (kinetic) energy to heat. There other examples of damping we could think about - here's
one:

Imagine hitting a cymbal. This causes the cymbal to oscillate. These oscillations cause the air around the
cymbal to vibrate, and these vibrations travel to your ear and you hear this as sound. The sound of the cymbal
will eventually die away, as the air resistance and internal losses within the cymbal reduce the restoring forces,
causing the oscillations to get smaller and smaller. Placing your hand on the cymbal after hitting it can greatly
speed up the damping process, as your fingers absorb the kinetic energy (being soft and a bit 'pudgy'!) very
effectively..

Normally it is hard to see a cymbal vibrating because it is moving too fast. Here we've slowed it down to 1/ 80th
of normal speed. You can see that the oscillations take ages to die away, as damping is small. For many
oscillations (including this one) the damping forces are roughly proportional to velocity, and this leads to an
exponential decay of amplitude over time.

Some drums use extra damping pads to control the time the skin rings for, after striking. Many drummers put a
pillow in the 'kick' (bass) drum - the extra damping creates more of a 'thud' than a 'boom'. The world of
acoustics and sound is full of examples where damping is intentionally used to control vibrations and sounds.

Some systems such as shock absorbers on cars or vibration dampers on bridges have a very high level of
damping. The controlled mass (car body, bridge deck etc) will ideally return to equilibrium in the shortest
possible time once vibrated, rather than bouncing around either side of it as we have seen up to now. Such
systems are said to be critically damped.
http://www.acoustics.salford.ac.uk/feschools/waves/shm2.php

Free Oscillations, Forced Oscillations and Resonance


If an oscillator is displaced and then released it will begin to vibrate. If no more external forces are applied to
the system it is a free oscillator. If a force is continually or repeatedly applied to keep the oscillation going, it is
a forced oscillator.

Let's use the example of a wine glass. If you 'flick' the glass and let it ring, the glass acts as a free oscillator. If
you wet your finger and run it around the glass rim, your finger will force continuous vibration by repeatedly
slipping and sticking (very fast). This slip-stick excitation is the same as takes place when violin players use a
violin bow to drive their strings.

In this extract from a slow motion video, a glass is driven into forced oscillation. Water was poured into the
glass, as it vibrates very visibly and makes the motion of the 'walls' of the glass very clear.

Going back to our example of a playground swing: if the swing is pushed just once it acts as a free oscillator
and the damping effects of air resistance and losses at the pivots mean it will eventually stop swinging. If the
swing is pushed each time it reaches a certain point it behaves as a forced oscillator and will continue to swing
for as long as energy is supplied. This is shown below:

The 'pusher' normally pushes the swing every time it reaches its maximum negative displacement. What would
happen if he pushed the swing more often, or less often?

If the swing is simply lifted and let go it will swing with a natural frequency. This frequency is determined by a
number of factors, of which the most important is the length of the swing from pivot to seat. How does this
natural frequency effect when the pusher should push?

If the swing is pushed with a long interval between pushes, the swing will not recieve enough energy to replace
that lost by damping. Also, the pushes may be 'out of synch' with the swing's natural period of oscillation. This
might mean that sometimes the pusher would be pushing when the swing is in the wrong location...

If the person pushes with a shorter interval between pushes, he may again be 'out of synch' with the natural
frequency of the swing..

http://www.acoustics.salford.ac.uk/feschools/waves/shm3.php

The natural frequency that the swing 'wants to' oscillate at is called its resonant frequency. If the pusher
pushes at the swing's resonant frequency, the amplitude of oscillation will build up. With each period, the
pusher will add more energy to the system. Eventually, what usually happens is that energy supply equals
energy loss (to damping), and the amplitude stabilises at some large value. If there is insufficient damping in
the system the oscillation amplitude can get very large and something dramatic may happen...

Here's a slow-motion video of a wine glass driven into forced oscillation at its resonant frequency. This time it's
forced by a high-power loudspeaker driven by a sine-wave generator. To vibrate, the glass walls must bend,
and glass won't bend too far before...

...it breaks! A little more damping (e.g. a rubber band around the glass) could stop this happening by reducing
the resonant vibration amplitude. If the oscillating system was the wing attached to the jet plane taking you on
holiday, you would be hoping that the designer had incorporated sufficient damping into the structure! Let's look
at that some more...

http://www.acoustics.salford.ac.uk/feschools/waves/shm3.php
More on damping
At resonance the amount of energy lost due to damping is equal to the rate of energy supply from the driver.
The driver is the source of external energy that keeps the oscillations going - for example, the person pushing a
kid on a swing. Increasing the damping will reduce the size (amplitude) of the oscillations at resonance, but the
amount of damping has next to no effect at all on the resonant frequency.

Damping also has an effect on the 'sharpness' of a resonance; sharpness is a not-very scientific way of
describing how sensitively the resonance is tuned, and is sometimes called the 'Q-factor' by engineers. If
damping is very small, a system will only oscillate a little if driven even slightly above or below the 'right'
frequency - but when the driver hits the resonant frequency 'bang-on', suddenly the oscillations can get very
large. Conversely, if damping is large, the amplitude of oscillations at resonance will decrease, but if the driver
shakes (excites) the system at the 'wrong' frequency, the system will still respond quite strongly. This means
there will be less of a resonant effect, but that it will happen over a larger range of frequencies.

The interactive animation below shows how the suspension of a car can be used to demonstrate resonance
and damping.

The 'sharpness' of a resonance is measured by its Q-Factor. Although this does have a precise mathematical
definition, it is most easily understood as roughly the number of free oscillations the oscillator will complete
before decaying to zero. At the M.O.T. garage, the mechanic tests the damping of your shock absorbers by
'bouncing' the wings of the car...he probably doesn't know he is measuring Q-factor, but he does know that
more than a 'bounce-and-a-half' and it fails!

So: a small amount of damping equates to a large Q, and a large amount of damping equates to a small Q.

http://www.acoustics.salford.ac.uk/feschools/waves/shm4.php

Simple Harmonic Motion (SHM)


Click here for questions & homework on SHM.

Click for SHM answers.

Objects can oscillate in all sorts of ways but a really important form of oscillation is
SHM or Simple Harmonic Motion.

An object is undergoing simple harmonic motion (SHM) if;

1. the acceleration of the object is directly proportional to its


displacement from its equilibrium position.
2. the acceleration is always directed towards the equilibrium position.
The frequency (f) of an oscillation is measure in hertz (Hz) it is the number of
oscillations per second. The time for one oscillation is called the period (T) it is
measured in seconds.

Acceleration we can calculate the acceleration of the object at any point in its
oscillation using the equation below.

In this equation; a = acceleration in ms-2, f = frequency in Hz, x = displacement from the


central position in m.

Displacement When using the equation below your calculator must be in radians
not degrees ! we can calculate the displacement of the object at any point in its
oscillation using the equation below.

The terms in this equation are the same as the equations above. The extra terms in this
equation are: A = the amplitude (maximum displacement) in m, t = the time since the
oscillation began in s.

Velocity we can calculate the velocity of the object at any point in its oscillation
using the equation below.
The terms in this equation are the same as the equations above. The extra term in this
equation is: v = the velocity in ms-1.

SHM graphs

When we plot the displacement, velocity and acceleration during SHM against time we
get the graphs below.

The velocity equation simplifies to the equation below when we just want to know the
maximum speed.

The acceleration equation simplifies to the equation below when we just want to know
the maximum acceleration.

Time period of a mass-spring system


Time period of a Pendulum

SHM and Energy

For a pendulum undergoing SHM energy is being transferred back and forth between
kinetic energy and potential energy. The total energy remains the same and is equal to
kinetic energy + potential energy (see graph below).
http://physicsnet.co.uk/a-level-physics-as-a2/further-mechanics/simple-harmonic-
motion-shm/

Basic Oscillations

The time taken for an oscillating object to complete one full oscillation is called the time
period, T. It is measured in seconds.

If a number of oscillations are involved we can work out the time period by dividing the total
time taken by the number of oscillations completed:

The frequency, f, of oscillations is the number of oscillations undergone in one second, and is
measured in hertz (Hz).

The frequency and the period can therefore be related as:

The displacement of an oscillating particle is the distance the particle has been moved from its
equilibrium position.

The amplitude of an oscillation is the maximum displacement of the vibrating object from the
equilibrium position (its usual position).

Note: Always check the x-axis on the graph, as it is easy to confuse wavelength and time period
on diagrams!
Simple Harmonic Motion

Simple Harmonic Motion (SHM) is a particular type of oscillation. It is useful because its time
period stays the same even when its amplitude changes. We'll come to the full definition later!

Lets think about a simple example of shm to work out the relationship between
displacement, velocity and acceleration:

Now remember that displacement, velocity and acceleration are all vectors, and as a result,
direction is important. Let's choose anything in the up-wards direction to
be positive, anything downwards to be negative. (If you decide to do the opposite, it doesn't
matter - just stick to your choice.)

If we set this system oscillating by lifting the mass and letting it go, then the system
starts with:

Maximum positive displacement (because it's above the middle).

Zero velocity (it's not moving at the first instant).

Maximum negative acceleration (because it is about to start moving down).

The interaction below shows how velocity and acceleration change in simple harmonic motion.
It shows the relationship between velocity and acceleration. Click "next" to see each part of
the motion...
The displacement, velocity and acceleration of the mass are related as shown above. To draw
these, think about what the object is doing at each point as it oscillates from the start position
described above.

As it passes through the equilibrium position on the way down it's at maximum speed down
(negative), its displacement is zero and because the spring is at its equilibrium position, there is
no resultant force on the mass so it is not accelerating.

At the bottom, the mass stops momentarily as it changes direction, so velocity is zero. The
displacement is a maximum in the negative direction, so the acceleration is a maximum in a
positive direction as the spring tries to shorten again.

The important point to note is the phase difference between these three variables...

1. The velocity, v, is zero where there are stationary points at the peaks and troughs of the
displacement graph and the velocity is a maximum when the displacement is zero. (Don't
forget the gradient of the displacement graph will equal velocity.)

2. The displacement and acceleration graphs are 180 degrees out of phase and therefore look
like a mirror image of each other in the time axis. (Don't forget the gradient of the velocity
graph will equal acceleration.)

Definition of Simple Harmonic Motion:

All of the above leads us to the formal definition of shm:

A body is undergoing SHMwhen the acceleration on the body is proportional to its


displacement, but acts in the opposite direction.

Acceleration a displacement

a-s

It's also important to note that for SHM, the time period of the oscillations is constant and
doesn't change even if the amplitude is changing.

There are two common examples of simple harmonic motion:


Where L = length of pendulum (m)
Where m = mass (kg)

g = acceleration due to gravity (ms-2)


and k = spring constant (Nm-1)

SHM is used to explain the behaviour of atoms in a lattice, which oscillate like masses on
springs.

http://www.s-cool.co.uk/a-level/physics/simple-harmonic-motion-and-
damping/revise-it/simple-harmonic-motion

Simple Harmonic Motion (SHM)

CUT TO THE CHASE

Introduction

In addition to linear motion and rotational motion there is another kind of motion that is

common in physics. This is the to and fro motion of oscilations or vibrations.

When something oscillates, it moves back and forth with time. It is helpful to trace out the

position of an oscillating particle with time so we can define some terminology.


Flash 1. Simple Harmonic Motion (SHM) of the position of a particle with time produces a
Sinusoidal wave.

PERIOD, AMPLITUDE AND FREQUENCY

The time taken for the particle to complete one oscilation, that is, the time taken for the

particle to move from its starting position and return to its original position is known as

the period. and is generally given the symbol T. The frequency is related to the period, it

is defined as how many oscillations occur in one second. Since the period is the time taken

for one oscillation, the frequency is given by

f = 1/T(1)

The frequency is measured in [s-1]. This unit is known as the Hertz (Hz) in honour of the

physicist Heinrich Hertz. The maximum displacement of the particle from its resting position

is known as the amplitude. The frequency is also given the symbol f.

Simple Harmonic Motion

The definition of simple harmonic motion is simply that the acceleration causing the

motion a of the particle or object is proportional and in opposition to its displacement x from

its equilibrium position.

a(t) -x(t)

Where k is a constant of proportionality. This remembering that the acceleration is the

second derivative of position, also leads us to the differential equation

x''(t) = - k x(t)

Simple Harmonic Motion is closely related to circular motion as can be seen if we take an

object that moves in a circular path, like a ball stuck on a turntable. If we consider just the

y-component of the motion the path with time we can see that it traces out a wave as

shown in Flash 2.

y(t) = A sin(t)(2)

Flash 2. Simple Harmonic Motion, is also a component of circular motion.


We can also see that the period of the motion is equal to the time it takes for one rotation.

Therefore, if we know the angular velocity = &theta/t. For one rotation, =

2/T therefore the period is also equal to

T = 2/(3)

The particle can also at different speeds is conected with the period. The frequency is the

number of oscilations per second.

Consider the particle undergoing simple harmonic motion in Flash 3. The displacement with

time takes the form of a sinusoidial wave. The velocity of the particle can be calculated by

differentiating the displacement. The result is also a wave but the maximum amplitude is

delayed, so that when the displacement is at a maximum the velocity is at a minimum and

when the displacement is zero the velocity has its greatest value.

Simple Harmonic Motion is characterised by the acceleration a being oppositely


proportional to the displacement, y.
Flash 3. Displacement, velocity and acceleration vectors of a particle undergoing simple
harmonic motion

We set this out mathematically, using a differential equation as in equation (4). We specify

the equation in terms of the forces acting on the object. The acceleration is the second

derivative of the position with respect to time and this is proportional to the position with

respect to time. The minus sign indicates that the position is in the opposite direction to the

acceleration.

m y''(t) = - k y(t)(4)

The derivation of the solution can be found here

For which the general solution is a wave like solution.

y(t) = c1 cos(t) + c2 sin(t)


Where, is the angular frequency. (=2f) The values of c1 and c2 are determined by the

initial conditions. Specifically, c1 = y0 and c2 = v0/ These two initial conditions specify the

starting position and the initial velocity.

The general solution can also be written more compactly as

y(t)= A cos(t - )(5)

Where = tan-1(y0/v0), A = (y02 + (v0/)2)1/2

Differentiating once with respect to time, we obtain the velocity. (The derivative of cos x = -

sin x)

v = y'(t)= - A sin(t - )(6)

Finally, the acceleration is the derivative of the velocity with respect to time. (The derivitive

of -sin x = - cos x)

a = y''(t) = - 2A cos(t - )(7)

Substituting equations (5) and (7) into equation (4) we verify that this does indeed satisfy

the equation for simple harmonic motion. With the constant of proportionality k = 2

Thus

a(t) = - 2y(t)

The time for the maximum velocity and acceleration can be determined from these

equations. From equation (6) the maximum magnitude of the velocity occurs when sin(t -

) is 1 or -1. Therefore the maximum velocity is A. Intuitively, we can imagine that this

velocity occurs when the oscillating system has reached the equilibrium position and is

about to overshoot. The minus sign indicates the direction of travel is in the opposite

direction.
The maximum acceleration occurs where the argument of cosine in eqn (7) is also -1 or 1.

Thus the maximum acceleration is 2A which occurs at the ends of the oscillations, as this

is where the direction changes.

Examples of SHM
SPRING MASS SYSTEM

Consider a spring of spring constant k conected to a mass m. If the mass is displaced from

its equilibrium position by a distance x a force F will act in the opposite direction to the

displacement. From Hooke's Law the magnitude of the force is given by

F= - k x

When the mass is released it the force will act on the mass to bring it back to its equilibrium

position. However, if there is no friction the inertia of the mass will cause it to overshoot the

equilibrium position and the force will act in the opposite direction slowing it down and

pulling it back.

The action of this force on the mass keeps it oscillating backwards and forwards.

Flash 4. Spring Mass system undergoing SHM.

The spring mass system consists of a spring with a spring constant of k attached to a

mass, m.The mass is displaced a distance x from its equilibrium position work is done and

potential energy is stored in the spring. If the mass is displaced by a small distance dx, the

work done in stretching the spring is given by dW = F dx.

The force on the spring is assumed to obey Hooke's law, therefore, the restoring force is

proportional to the extension. The work done is then dW = - k x dx


Making the total work W = - k x dx = -1/2 k x2 + C

At the time of release, the energy of the system will consist totally of potential energy. PE=

-1/2 kx2 and the potential energy as a function of time is


PE(t) = 1/2k A cos2(t - )

As the spring pulls the mass toward the equilibrium position, the potential energy is

transformed into kinetic energy until at the equilibrium position the kinetic energy will be

maximised. The KE is 1/2 mv2 and the maximum velocity occurs at x=0. From equation (6),

the velocity will be A

Therefore the kinetic energy is 1/2 m2A2 sin2(t - )

From the conservation of mechanical energy, the sum of energy between kinetic energy and

potential energy will always be the same.

U = KE(t) + PE(t) = 1/2 m2A2 sin2(t - ) + 1/2k A cos2(t - )

1/2m2x2 = 1/2k x2

is transferred into kinetic energy by moving the mass. For a particle undergoing simple

harmonic motion, the displacement x is given by equation 1. The potential energy is

gradually transferred to kinetic energy. Kinetic energy is given by 1/2 mv2. The velocity is

given by equation (6). Summing the kinetic energy and potential energy we obtain,

U= (1/2)A2( k + m)

= 1/2 k A2(8)

Since k = m2 and cos2x + sin2x = 1

Flash 5 shows the change in energy betweeen kinetic energy and potential energy with

time. The blue line shows the potential energy. It is highest at the positions of maximum

displacement. The red line shows the kinetic energy. It has a maximum when the velocity is

greatest, ie. as it passes the equilibrium position. The green line shows the total mechanical

energy of the system, ie. the sum of the potential energy and kinetic energy. The total

energy remains constant because there are no losses to friction, heat or air resistance.

Flash 5. KE (red) and PE (blue) curves. Total energy (green) remains constant
SIMPLE PENDULUM

Simple Pendulum

Another common example used to illustrate simple harmonic motion is the simple

pendulum. This idealised system has a one end massless string suspended a mass m and

the other end fixed to a stationary point. If the mass is displaced by a small distance, the

angle moved is small.

The torque on the fixed point P is = I

- mg sin (t)L = mL2 ''(t)

''(t) +g/L (t) = 0

This has the same form as simple harmonic motion equation, x''(t) - 2x(t), and so the

solution is (t) = 0 cos(t - &phi)

the angular frequency is = (g/L)1/2.


It is interesting to note that the mass does not appear in this equation. This means that the

frequency of the period only depends on the length of the string and the force of gravity.

Pendulums with shorter strings will oscillate faster than pendulums with longer strings. And

the same pendulum on the moon, where the force of gravity is 1/6th that of the gravity on

the Earth, will also take longer to oscillate.

We have glossed over one important aspect, in that this analysis is true only for small

angles of theta. We had to make the approximation that sin is aproximately the same as

which is true only for small angles. The real differential equation

''(t) +g/L sin (t) = 0

is non-linear and cannot be solved analytically.

TORSIONAL PENDULUM

Figure. A torsional pendulum.

A mass suspended to a fixed support by a thin wire can be made to twist about its axis. This

is known as a torsional pendulum. The mass attached to the wire rotates in the horizontal

plane. In this case is the angle of rotation. When the wire is untwisted and in equilibrium,

the angle is 0 degrees. It is the twisting of the wire that creates a restoring torque due to

the resistance of the wire to the deformation. For small angles of the magnitude of the

torque is proportional to the angle

=-k
Where k is the torque constant of the wire. As with the simple pendulum the equation of

motion is

= I

Where I is the moment of inertia

k(t) = - I''(t)

Once again we have formed the equation for simple harmonic motion and can write the

solution as (t) = A cos(t - )

The angular frequency, is given by

= (k/I)1/2

Summary

An oscillation follows simple harmonic motion if it fulfils the following two rules:

Acceleration is always in the opposite direction to the displacement from the


equilibrium position

Acceleration is proportional to the displacement from the equilibrium position

The acceleration and displacement are linked by the following equation:

a(t) = - 2 x(t)

Here is called the angular frequency of oscillation, and is given by 2 /T or 2 f

T is the period of oscillation (s), f=1/T = frequency of oscillation (Hz) and x is the

displacement (m).

Using Newton's 2nd Law (F=ma) we can show that the Force on the object due to inertia

will be:

F= -m2 x
Using Hooke's Law for springs (F=-kx), we know that the force on the oscillating mass due

to the springs is simplyF=-kx

These two forces are always in balance, so m 2 x - kx = 0

From this we can find the resonant frequency:

2=(kx) / (mx) = k/m,

= (k/m)1/2

f = 1/ 2 * [(k/m)1/2]

http://www.splung.com/content/sid/2/page/shm

Simple Harmonic Motion Concepts


INTRODUCTION
Have you ever wondered why a grandfather clock keeps accurate time? The motion of the pendulum is
a particular kind of repetitive or periodic motion called simple harmonic motion, or SHM. The position of
the oscillating object varies sinusoidally with time. Many objects oscillate back and forth. The motion of
a child on a swing can be approximated to be sinusoidal and can therefore be considered as simple
harmonic motion. Some complicated motions like turbulent water waves are not considered simple
harmonic motion.When an object is in simple harmonic motion, the rate at which it oscillates back and
forth as well as its position with respect to time can be easily determined. In this lab, you will analyze a
simple pendulum and a spring-mass system, both of which exhibit simple harmonic motion.

DISCUSSION OF PRINCIPLES
A particle that vibrates vertically in simple harmonic motion moves up and down between two
extremes y = A. The maximum displacement A is called the amplitude. This motion is shown
graphically in the position-versus-time plot in Figure 1.
Figure 1: Position plot showing sinusoidal motion of an object in SHM

One complete oscillation or cycle or vibration is the motion from, for example,

y = A

to

y = +A

and back again to

y = A.

The time interval T required to complete one oscillation is called the period. A related quantity is
the frequency f, which is the number of vibrations the system makes per unit of time. The
frequency is the reciprocal of the period and is measured in units of Hertz, abbreviated Hz;

1 Hz = 1 s1.

(1)
f = 1/T
If a particle is oscillating along the y-axis, its location on the y-axis at any given instant of time t,
measured from the start of the oscillation is given by the equation

(2)
y = A sin(2ft).
Recall that the velocity of the object is the first derivative and the acceleration the second derivative of
the displacement function with respect to time. The velocity v and the acceleration a of the particle at
time t are given by the following.

(3)
v = 2fA cos(2ft)
(4)
a = (2f)2[A sin(2ft)]
Notice that the velocity and acceleration are also sinusoidal. However, the velocity function has a 90
or /2 phase difference while the acceleration function has a 180 or phase difference relative to the
displacement function. For example, when the displacement is positive maximum, the velocity is zero
and the acceleration is negative maximum.Substituting from equation 2 into equation 4 yields

(5)
a = 42f2y.
From equation 5, we see that the acceleration of an object in SHM is proportional to the displacement
and opposite in sign. This is a basic property of any object undergoing simple harmonic
motion.Consider several critical points in a cycle as in the case of a spring-mass system in oscillation. A
spring-mass system consists of a mass attached to the end of a spring that is suspended from a stand.
The mass is pulled down by a small amount and released to make the spring and mass oscillate in the
vertical plane. Figure 2 shows five critical points as the mass on a spring goes through a complete
cycle. The equilibrium position for a spring-mass system is the position of the mass when the spring is
neither stretched nor compressed.
Figure 2: Five key points of a mass oscillating on a spring

The mass completes an entire cycle as it goes from position A to position E. A description of each
position is as follows.

Position A: The spring is compressed; the mass is above the equilibrium point at

y=A

and is about to be released.Position B: The mass is in downward motion as it passes through the
equilibrium point.Position C: The mass is momentarily at rest at the lowest point before starting on its
upward motion.Position D: The mass is in upward motion as it passes through the equilibrium
point.Position E: The mass is momentarily at rest at the highest point before starting back down again.

By noting the time when the negative maximum, positive maximum, and zero values occur for the
oscillating object's position, velocity and acceleration, you can graph the sine (or cosine) function. This
is done for the case of the oscillating spring-mass system in the table below and the three functions
are shown in Figure 3. Note that the positive direction is typically chosen to be the direction that the
spring is stretched. Therefore, the positive direction in this case is down and the initial position A in
Figure 2 is actually a negative value. The most difficult parameter to analyze is the acceleration. It
helps to use Newton's second law, which tells us that a negative maximum acceleration occurs when
the net force is negative maximum, a positive maximum acceleration occurs when the net force is
positive maximum and the acceleration is zero when the net force is zero.
Figure 3: Position, velocity and acceleration vs. time

For this particular initial condition (starting position at A in Figure 2), the position curve is a cosine
function (actually a negative cosine function), the velocity curve is a sine function, and the
acceleration curve is just the negative of the position curve.

Mass and Spring


A mass suspended at the end of a spring will stretch the spring by some distance y. The force with
which the spring pulls upward on the mass is given by Hooke's law

(6)
F = ky
where k is the spring constant and y is the stretch in the spring when a force F is applied to the
spring. The spring constant k is a measure of the stiffness of the spring.The spring constant can be
determined experimentally by allowing the mass to hang motionless on the spring and then adding
additional mass and recording the additional spring stretch as shown below. In Figure 4a, the
weight hanger is suspended from the end of the spring. In Figure 4b, an additional mass has been
added to the hanger and the spring is now extended by an amount

y.

This experimental set-up is also shown in the photograph of the apparatus in Figure 5.
Figure 4: Set up for determining spring constant

Figure 5: Photo of experimental set-up

When the mass is motionless, its acceleration is zero. According to Newton's second law, the net force
must therefore be zero. There are two forces acting on the mass: the downward gravitational force and
the upward spring force. See the free-body diagram in Figure 6 below.

Figure 6: Free-body diagram for the spring-mass system

So Newton's second law gives us

(7)
mg ky = 0
where

is the change in mass and

is the change in the stretch of the spring caused by the change in mass, g is the gravitational
acceleration, and k is the spring constant. Equation 7 can also be expressed as

(8)
m =

k
g
y.
Newton's second law applied to this system is

ma = F = ky.

Substitute from equation 5 for the acceleration to get

(9)
m(42f2y) = ky
from which we get an expression for the frequency f and the period T.

( 10 )
f=

1
2
k
m

( 11 )
T = 2

m
k

Using equation 11, we can predict the period if we know the mass on the spring and the spring
constant. Alternately, knowing the mass on the spring and experimentally measuring the period, we
can determine the spring constant of the spring.Notice that in equation 11, the relationship
between T and m is not linear. A graph of the period versus the mass will not be a straight line. If we
square both sides of equation 11, we get
( 12 )
T2 = 42
m
k
.
Now, a graph of

T2

versus m will be a straight line, and the spring constant can be determined from the slope.

Simple Pendulum
The other example of simple harmonic motion that you will investigate is the simple pendulum. The
simple pendulum consists of a mass m, called the pendulum bob, attached to the end of a string.
The length L of the simple pendulum is measured from the point of suspension of the string to the
center of the bob as shown in Figure 7 below.

Figure 7: Experimental set-up for a simple pendulum

If the bob is moved away from the rest position through some angle of displacement as in Figure 8,
the restoring force will return the bob back to the equilibrium position. The forces acting on the bob are
the force of gravity and the tension force of the string. The tension force of the string is balanced by
the component of the gravitational force that is in line with the string (i.e. perpendicular to the motion
of the bob). The restoring force here is the tangential component of the gravitational force.
Figure 8: Simple pendulum

When we apply trigonometry to the smaller triangle in Figure 8, we get the magnitude of the restoring
force |F| = mg sin . This force depends on the mass of the bob, the acceleration due to gravity g, and
the sine of the angle through which the string has been pulled. Again Newton's second law must apply,
so

( 13 )
ma = F = mg sin
where the negative sign implies that the restoring force acts opposite to the direction of motion of
the bob.Since the bob is moving along the arc of a circle, the angular acceleration is given by

= a/L.

From equation 13 we get

( 14 )
=

g
L
sin .
In Figure 9, the blue solid line is a plot of sin() versus , and the straight line is a plot of in degrees
versus in radians. For small angles, these two curves are almost indistinguishable. Therefore, as long
as the displacement is small, we can use the approximation sin .

Figure 9: Graphs of sin versus

With this approximation, equation 14 becomes

( 15 )
=

g
L
.
Equation 15 shows the (angular) acceleration to be proportional to the negative of the (angular)
displacement, and therefore the motion of the bob is simple harmonic and we can apply equation 5 to
get

( 16 )
= 42f2 .
Combining equation 15 and equation 16 and simplifying, we get

( 17 )
f=

1
2
g
L

and
( 18 )
T = 2

L
g

.
Note that the frequency and period of the simple pendulum do not depend on the mass.

Copyright 2013 Advanced Instructional Systems, Inc. and North Carolina State University | Credits

http://webassign.net/question_assets/ncsucalcphysmechl3/lab_7_1/manual.html

Simple Harmonic Motion

page 1 of 2

Having established the basics of oscillations, we now turn to the special case
of simple harmonic motion. We will describe the conditions of a simple
harmonic oscillator, derive its resultant motion, and finally derive the energy of
such a system.

The Simple Harmonic Oscillator

Of all the different types of oscillating systems, the simplest, mathematically


speaking, is that of harmonic oscillations. The motion of such systems can be
described using sine and cosine functions, as we shall derive later. For now,
however, we simply define simple harmonic motion, and describe the force
involved in such oscillation.

To develop the idea of a harmonic oscillator we will use the most common
example of harmonic oscillation: a mass on a spring. For a given spring with
constant k , the spring always puts a force on the mass to return it to the
equilibrium position. Recall also that the magnitude of this force is always
given by:

F(x) = - kx
where the equilibrium point is denoted by x = 0 . In other words, the more the spring is stretched
or compressed, the harder the spring pushes to return the block to its equilibrium position. This
equation is only valid if there are no other forces acting on the block. If there is friction between
the block and the ground, or air resistance, the motion is not simple harmonic, and the force on
the block cannot be described by the above equation.

Though the spring is the most common example of simple harmonic motion, a
pendulum can be approximated by simple harmonic motion, and the torsional
oscillator obeys simple harmonic motion. Both of these examples will be
examined in depth in Applications of Simple Harmonic Motion.

Simple Harmonic Motion

>From our concept of a simple harmonic oscillator we can derive rules for the
motion of such a system. We start with our basic force formula, F = - kx .
Using Newton's Second Law, we can substitute for force in terms of
acceleration:

ma = - kx
Here we have a direct relation between position and acceleration. For you calculus types, the
above equation is a differential equation, and can be solved quite easily. Note: The following
derivation is not important for a non- calculus based course, but allows us to fully describe the
motion of a simple harmonic oscillator.

Deriving the Equation for Simple Harmonic Motion

Rearranging our equation in terms of derivatives, we see that:

m = - kx

or

+ x=0
Let us interpret this equation. The second derivative of a function of x plus the function itself
(times a constant) is equal to zero. Thus the second derivative of our function must have the
same form as the function itself. What readily comes to mind is the sine and cosine function. Let
us come up with a trial solution to our differential equation, and see if it works.
http://www.sparknotes.com/physics/oscillations/oscillationsandsimpleharmonicmotio
n/section2.rhtml

Simple Harmonic Motion

page 2 of 2

As a tentative solution, we write:

x = a cos(bt)
where a and b are constants. Differentiating this equation, we see that

= - ab sin(bt)

and

= - ab 2cos(bt)
Plugging this into our original differential equation, we see that:

- ab 2cos(bt) + a cos(bt) = 0

It is clear that, if b 2 = , then the equation is satisfied. Thus the equation


governing simple harmonic oscillation is:

simple

x = a cos t
The Equation for Simple Harmonic Motion

From the equation for simple harmonic motion we can tell a lot about the motion of a
harmonic system. First of all, x is maximum when the cosine function is equal to 1, or
when x = a . Thus a in this equation is the amplitude of oscillation, which we have
already denoted by x m . Secondly, we can find the period of oscillation of the system.
At t = 0 , x = x m . Also, at t = 2 , x = x m . Since both these instances have the
same position, the time between the two gives us our period of oscillation. Thus:

T = 2

and

= =

finally,

= 2 =

Note that the values of period and frequency depend only on the mass of the block
and the spring constant. No matter what initial displacement is given to the block, it
will oscillate at the same frequency. This concept is important. A block with a small
displacement will move with slower velocity, but with the same frequency as a block
with a large displacement.

Notice also that our value for is the same as what we called the constant b in our
original equation. So now we know that a = x mand b = . In addition we can take the
time derivative of our equation to generate a full set of equations for simple harmonic
motion:

x=x mcos(t)
v=- x msin(t)
2
a=- x mcos(t)

Thus we have derived equations for the motion of a given simple harmonic system.

Energy of a Simple Harmonic Oscillator

Consider a simple harmonic oscillator completing one cycle. In the jargon of


conservative vs. nonconservative forces (see Conservation of Energy the oscillator has
completed a closed loop, and returns to its initial position with the same energy it
began with. Thus the simple harmonic oscillator is a conservative system. Since the
velocity of the oscillator does change, however, there must be an expression for the
potential energy of the system, such that the total energy of the system is constant.

We already know the kinetic energy of the system at any given time:

2
K= mv

= m(- x msin(t))2

2
= kx m sin2(t)

The kinetic energy has a maximum value when the potential energy is zero,
and sin(t) = 1 . Thus K max = kx m . Since the potential energy is zero at this
point, this value must give the total energy of the system. Thus, at any time, we can
state that:
E=U + K

2 2
kx m =U + kx m sin2(t)

Solving for U:

2
U= kx m (1 - sin2(t))

Recall that sin2 a + cos2 a = 1 . We can thus substitute:

2
U= kx m cos2(t)
However, we also know that x = x mcos(t) for any simple harmonic oscillation.
Using this knowledge we can further simplify our equation for potential energy:

simplify

2
U= kx

With this equation we have an expression for the potential energy of a simple
harmonic oscillator given a displacement from equilibrium. When examined
practically, this equation makes sense. Consider our example of a spring. When the
spring is stretched or compressed a large amount (i.e. the block on the spring has a
large magnitude for x ), there is a great deal of energy stored in those springs. As
the spring relaxes and accelerates the block this potential energy gets converted to
kinetic energy. Shown below are three positions of the oscillating spring, and the
energies associated with each position.
Figure %: An oscillating block a) at equilibrium with only kinetic energy b) at
maximum displacement with only potential energy c) partially displaced with both
2
potential and kinetic energy. Note that the total energy in each case is kx m .

This SparkNote introducing oscillation and simple harmonic motion involved a great
deal of mathematics and theoretical calculations. In the next SparkNote we explore
oscillations on a more practical level, examining real physical situations and various
types of oscillators.

http://www.sparknotes.com/physics/oscillations/oscillationsandsimpleharmonicmotio
n/section2/page/2/

S-ar putea să vă placă și