Phys 212 Lecture Notes Part 1

Download as pdf or txt
Download as pdf or txt
You are on page 1of 96

PHYS 212 VIBRATIONS AND WAVES PART 1

LECTURE NOTES

Lecture 1 Periodic Motion and Simple Harmonic Oscillation

These first few lectures are meant to accompany and supplement Chapter 15 of
Benson’s University Physics. In particular, you should read Sections 15.1–15.4
along with these notes, and you should be able to work through the associated
exercises and problems in the text by the end of Lecture 4.

A periodic phenomenon is one that repeats at regular intervals. Just a few examples of such
periodic events are the vibration of a string; the alternating compression and rarefaction of
air in a sound wave; or the orbit of a planet about a star. As these examples suggest, we
are often concerned with phenomena that are periodic in time, in which case we use the term
period to mean the separation in time between consecutive repetitions of an event (say, the
closest approach of a planet to the star it orbits) or consecutive realizations of a given state of
a physical system (e.g., every time all of the position, velocity, and acceleration of a pendulum
return to some previously attained values). We normally use an upper-case T to denote the
period.

An oscillation is a periodic fluctuation in the value of a physical quantity between two ex-
tremes, above and below some intermediate (equilibrium) value. There are two basic types
of oscillation. The first type is mechanical oscillation, which involves the bodily motion
of physical objects in space (for example, the oscillations of atoms about their equilibrium
positions in a crystal lattice; the swinging of a pendulum; or the distance of a planet from
any fixed point). There are also non-mechnical oscillations, involving periodic variations in
the values of physical properties, rather than spatial displacements explicitly. Some examples
are the oscillation of density and pressure in a sound wave, the oscillation of voltage in an
alternating-current (AC) electrical circuit, and the oscillation of electric- and magnetic-field
strengths in a light wave.

We begin here by focusing on mechanical oscillations: What are the main features of such
oscillations? How do we describe them and analyze them mathematically? How do they
apply to real physical systems? But the main physical concepts and the basic mathematical
techniques that we develop in the context of mechanical oscillations apply equally to non-
mechanical oscillations. Then, oscillations of any kind lead naturally to the idea of waves,
as travelling or propagating patterns of oscillatory disturbances. Energy is transported by
waves, and energy plays a central role in our physical understanding of the world and Universe
around us. As such, an understanding of waves is key to many processes and phenomena in
various areas of physics—from acoustics (the study of sound) and optics (light), through to
mechanics and dynamics (with engineering applications), electromagnetism (light again; but
also other forms of radiation) and so on, even up to to much more “modern” fields like general
relativity (gravity) and quantum mechanics (atomic- and subatomic-scale physics). We will
–2–

touch specifically on quantum mechanics at the end of this module, when we discuss the famous
Schrödinger equation.

To start, let us consider a particularly simple example of periodic motion; that of a point
(a particle, a body, an object, . . . it doesn’t matter what, exactly, at this level) traversing a
circular path about the origin in the (x, y) plane. A snapshot of this motion at some unspecified
time t is illustrated in Figure 1. Let us assume that the motion is in the counterclockwise
direction at constant speed. The “orbiting” point is always a distance A from the origin of
our (x, y) coordinate system, and thus A is the radius of the circle shown. The line joining
the origin to the instantaneous position of the point makes a time-varying angle φ(t) to the
x-axis in this set-up. We refer to φ(t) as the phase. With this angle measured in radians, the
arc-length s measuring the distance along the circle from the positive x-axis to the point at
time t is
s(t) = Aφ(t) .
Here A has the usual units of length, and thus so does s(t). If the speed ds/dt is constant,
then the fact that A is constant immediately implies that the time-derivative of the phase is

Fig. 1.— Coordinates for a body of mass m moving on a circular path of radius A about the
origin. The arc length s(t) = Aφ(t) for φ measured in radians.
–3–

also constant:
dφ 1 ds
ω≡ = = constant .
dt A dt
As a result, the phase depends on time simply as
Z Z Z
φ(t) = dφ = ωdt = ω dt = ωt + φ0 , (1)

where φ0 is a constant of integration—called the phase constant—which is related to the


initial conditions of a particular system. That is, φ = φ0 at time t = 0; so the phase constant
keeps track of the state of the system when we decide to start measuring or observing it. Its
value is therefore dependent on when we choose to call t = 0. It is important always to know
what the phase constant is in any specific problem—it need not be 0, and so in general it
cannot be dropped from either algebraic or numerical calculations—but φ 0 is not usually of
much physical interest beyond its role as a “zeropoint.”

it is clear that the circular motion illustrated in Figure 1 is periodic; but to be rigorous about
it, suppose that the phase φ(t) is known for some specific time t. Then for a later time (t + T ),
if φ has increased by exactly 2π radians then the point has returned to exactly the same (x, y)
position that it had at time t. This identifies the time interval T as the period in this example;
and, from equation (1),

φ(t + T ) = φ(t) + 2π
=⇒ ω(t + T ) + φ0 = ωt + φ0 + 2π
=⇒ ωt + ωT + φ0 = ωt + φ0 + 2π

=⇒ T = . (2)
ω
[In the first line, φ(t + T ) is the function φ, evaluated at time t + T . In the second line, ω(t + T )
is the multiplication ω × (t + T ), from the definition of the function φ in equation (1).]

We have inferred the expression in equation (2) relating T and ω by considering circular motion
(from which it should already be familiar), but in fact the result applies much more generally
in the context of oscillations. In general, for oscillatory motion with a period T , we can always
talk about a quantity

ω= , (3)
T
which is called the angular frequency. The “frequency” part of this name reflects the fact
that ω ∝ 1/T , where 1/T = f is the usual frequency of periodic motion, in cycles per second.
The “angular” part of the name for ω comes from multiplying f by 2π radians/cycle. As we
have seen, ω is the rate of change of the phase, φ. If time is measured in seconds, then the the
units of T are obviously seconds, and ω therefore has units of s−1 (it could be thought of as
radians per second, but radians are strictly dimensionless numbers and so have no real units
at all).

Now consider the separate x- and y-components of the motion illustrated in Figure 1. From
the diagram,

x(t) = A cos(ωt + φ0 )
and y(t) = A sin(ωt + φ0 ) , (4)
–4–

so that each of x and y varies between extremes of +A and −A (since the maximum and
minimum values of cos and sin are +1 and −1), with a period of T = 2π/ω (this is the time
to go from +A to −A and back again). The averages of x and y over one full period are
hxi = hyi = 0. Thus, either of equations (4) describes an oscillation of amplitude A (which
is constant) around an average displacement of 0.

Still considering x and y separately, it is instructive (even outside of applications involving


circular motion, an analogy that we shall soon leave behind) to consider the first two time
derivatives of the displacements. What are the velocity and acceleration as functions of time,
in each of the x and y directions?

First, differentiate x once with respect to time:


dx d d
≡ ẋ = [A cos(ωt + φ0 )] = A × [cos(ωt + φ0 )] .
dt dt dt
Here we have introduced the notation ẋ as convenient shorthand for dx/dt, i.e., a single dot
over a quantity denotes the first time derivative of that quantity. The amplitude A
has come out of the derivative on the right-hand side because A is a constant. To evaluate
what’s left, we need to find the derivative of a function (cos) of a function (ωt + φ 0 ) of time,
which means that we need the chain rule. Returning to the fact that φ(t) = ωt + φ 0 , we write

d d dφ
ẋ = A × cos[φ(t)] = A × (cos φ) × = A × (− sin φ) × (ω) .
dt dφ dt
Thus, rearranging slightly and substituting back in the full expression for φ as a function of
time—and then using the chain rule again to differentiate y as well, with respect to time—
equation (4) ultimately gives

dx
≡ ẋ = −ωA sin(ωt + φ0 )
dt
dy
and ≡ ẏ = +ωA cos(ωt + φ0 ) . (5)
dt
Therefore, the velocities in the separate x and y directions also oscillate, about averages of
hẋi = hẏi = 0, with the same period T = 2π/ω as the x and y displacements, but now between
extremes of +ωA and −ωA.

Differentiating once more, using the chain rule in the same way, we find for the accelerations,

d2 x
≡ ẍ = −ω 2 A cos(ωt + φ0 )
dt2
d2 y
and ≡ ÿ = −ω 2 A sin(ωt + φ0 ) . (6)
dt2
Once again, the accelerations oscillate about averages of hẍi = hÿi = 0 with period T = 2π/ω.
The amplitude of the oscillation in acceleration is ω 2 A.

In equation (6), ẍ has been introduced to represent d2 x/dt2 , and ÿ stands for d2 y/dt2 . In
general, two side-by-side dots over a quantity denote the second time derivative of
that quantity.
–5–

Putting equations (4), (5), and (6) all together now:

x = A cos(ωt + φ0 ) y = A sin(ωt + φ0 )
dx dy
≡ ẋ = −ωA sin(ωt + φ0 ) ≡ ẏ = +ωA cos(ωt + φ0 ) (7)
dt dt
d2 x d2 y
≡ ẍ = −ω 2 A cos(ωt + φ0 ) ≡ ÿ = −ω 2 A sin(ωt + φ0 )
dt2 dt2
Some aspects of these two parallel sets of equations are worth commenting on in detail:

1. The only difference between the x and y equations is that the latter involve sin when
the former involve cos, and vice versa (plus a sign difference in the velocities). However,
since sin[(ωt + φ0 ) + π/2] = cos(ωt + φ0 ) and cos[(ωt + φ0 ) + π/2] = − sin(ωt + φ0 ),
this difference between x and y amounts to nothing more than a constant difference
of π/2 radians (= 90◦ , or one-quarter of a cycle) in their phases, φ = (ωt + φ0 ). Put
another way, we could still write y = A sin(ωt + φ0 ), as we have, but then also write
x = A sin(ωt + Φ0 ) with a different phase constant Φ0 = φ0 + π/2—in which case we
would find ẋ = +ωA cos(ωt + Φ0 ) and ẍ = −ω 2 A sin(ωt + Φ0 ). (Use the relations for sin
and cos of compound angles in the Maths and Physics Handbook to convince yourself
that these equations are equivalent to the x-equations just above when Φ0 = φ0 + π/2
as just stated.) Thus, we can choose to write either x or y in terms of either sin or cos
without affecting the basic behaviour of the displacements, velocities, and accelerations
(or the behaviour of any other physical quantities derived from these) as functions of
time; whatever we choose, we just have to be self-consistent, and use the correct phase
constant.

2. If there are no forces acting on the orbiting particle in Figure 1 other than the centripetal
force maintaining the circular motion in the first place, then the total energy of the
particle is just its kinetic energy. If the particle mass is m, then
1 1 ¡ 2
Etot = mv 2 = m ẋ + ẏ 2
¢
2 2
1 £ 2 2 2
m ω A sin (ωt + φ0 ) + ω 2 A2 cos2 (ωt + φ0 )
¤
=
2
1
mω 2 A2 sin2 (ωt + φ0 ) + cos2 (ωt + φ0 )
£ ¤
=
2
1
= mω 2 A2 , (8)
2
in which the last line follows from the fact that sin2 θ + cos2 θ = 1 for any angle θ.
Although we have derived it specifically from considerations of circular motion, the result
that the total energy is proportional to the square of the amplitude of oscillation is much
more generally applicable—as we shall see in future lectures. Note also that, since we
have stipulated that both the angular frequency ω and the amplitude A are constant
in time, in order to arrive at equations (7), the total energy Etot is also constant in
time in the current example. (Again, this is in the absence of additional forces, so this
conservation of energy is not at all unexpected.)

3. Just as the sin-versus-cos difference between x and y means that these displacements
are out of phase by π/2 rad, or 90◦ , or one-quarter of a cycle, so too the difference
–6–

between sin and cos in x versus ẋ (or the similar difference between y and ẏ) means
that displacement and velocity are π/2 rad out of phase (though they also have different
amplitudes). That is, if we continue to write x = A cos(ωt + φ0 ), then we could equally
well manipulate ẋ = −ωA sin(ωt + φ0 ) to write ẋ = +ωA cos(ωt + φ0 + π/2) instead. It
can then be seen that the phase of velocity (i.e., the argument of the cosine) is always
different from the phase of displacement by exactly π/2 rad. This result is again not
restricted to circular-motion situations, but applies more generally to a large class of
oscillatory motions.
In practical terms, being 90◦ out of phase means that when x is at a maximum (x = +A,
which happens for ωt + φ0 = 0, 2π, 4π, . . .), then ẋ = 0 [since sin(2nπ) = 0 for all integers
n]. Also when x is a minimum (x = −A, for ωt + φ0 = π, 3π, 5π, . . .), then ẋ = 0.
Conversely, when x is at its equilibrium or average value of x = 0 (which happens for
ωt + φ0 = π/2, 3π/2, 5π/2, 7π/2, . . .), then the velocity is at a minimum or maximum,
ẋ = ±ωA.

4. Similarly, the acceleration ẍ is a further π/2 rad out of phase with the velocity ẋ, and
thus ẍ is a full π radians (= 180◦ , or half a cycle) out of phase with the displacement x.
Therefore, when the velocity is at a maximum or minimum, the acceleration is 0 (as is the
displacement itself); and when the velocity is 0, the acceleration is at one of the extremes
±ω 2 A (and the displacement is at one of its extremes, |x| = A). More specifically still,
equations (7) show immediately that when x is at its maximum of +A, the acceleration
is at its minimum ẍ = −ω 2 A; and when the displacement is at the minimum x = −A,
the acceleration is maximized, ẍ = +ω 2 A. All of this applies as well to ÿ, ẏ, and y.

This last point is particularly important, as it illustrates the fact that, for motion of this
kind (again, in either x or y considered separately), the acceleration is proportional in
magnitude and opposite in direction to the displacement. Mathematically, we have
shown that if we begin with either of the two equations (4) for displacement as a function of
time—so, again, with

x(t) = A cos(ωt + φ0 )
or y(t) = A sin(ωt + φ0 )

—then, by comparing the first and third lines of equations (7), we infer accelerations (and
thence forces) related to displacement by

d2 x
ẍ = −ω 2 x ⇐⇒ + ω2x = 0
dt2
d2 y
or ÿ = −ω 2 y ⇐⇒ + ω2y = 0 .
dt2
Note that these last equations involving x and y are now completely identical, aside from the
fact that the displacement is labelled as x in one and as y in the other. Again, then, the cos and
sin forms that we have been working with for x and y are effectively interchangeable and lead
to a single equation for acceleration as a function of displacement. Note also that neither the
constant amplitude A nor the phase constant φ0 —both of which appear in the full expression
for x(t) or y(t)—appears in the last equations. This fact is typical of such equations—known
–7–

as second-order, ordinary differential equations—which involve the second derivative of


a physical variable.

The detailed discussion above has been meant to introduce some of the basic concepts and ter-
minology that arise when dealing with a very general class of physical systems, which undergo
simple harmonic oscillation, or SHO (also referred to as simple harmonic motion, or
SHM, when the oscillations involved are mechanical osillations). When the oscillation is one-
dimensional (in, say, the x-direction), or when it involves some non-mechanical scalar property
(such as density or energy, say; but we still denote the unspecified property by x for now),
then even if there is no explicit connection to any kind of circular motion, SHO or SHM is
defined by a differential equation, linking the second time derivative ẍ to the displacement
from an average (or equilibrium) x = 0 via

d2 x
ẍ ≡ = −ω 2 x for Simple Harmonic Oscillation. (9)
dt2
As we have shown explicitly, this equation is satisfied when the quantity x varies sinusoidally
with time:
x(t) = A sin(ωt + φ0 ) . (10)
Equivalently, we say that equation (10) is the solution to equation (9). As discussed above, we
could have written the solution as x(t) = A cos(ωt + φ0 ) for a different choice of φ0 , but from
now on we will adopt the sin form for compatibility with Benson’s University Physics.

By analogy with our dicussion starting from circular motion—but even in more general situa-
tions now, when there is no explicit connection to circular motion—the solution (10) for x(t)
in SHO is characterized by the following quantities:

ω the angular frequency of the oscillation


T = 2π/ω the period of the oscillation
A the amplitude of the oscillation
φ0 the phase constant (reflecting initial conditions)
φ = ωt + φ0 the phase of the oscillation (varies in time)

Additionally, the familiar frequency of cycles per second is

f = 1/T = ω/2π .

Finally, in a slightly more special case—but one still not restricted just to circular motion—if
the oscillation is mechanical, such that x is the spatial displacement from equilibrium of a body
with mass m and the amplitude A has units of length, then the total energy of the oscillator is
1
Etot = mω 2 A2 .
2

In subsequent lectures, we will be looking more closely at various physical quantities and
systems that are governed by the basic equation (9) and its solution (10), drawing on knowledge
of the main features that we have discussed here.
Lecture 2 Simple Harmonic Motion

1. More on Periodic Motion

In Lecture 1 we saw that a quantity varying in time as

x(t) = A sin(ωt + φ0 ) (1)

satisfies the second-order, ordinary differential equation

d2 x
ẍ ≡ = −ω 2 x (2)
dt2
relating the second time derivative of x to x itself. Either of these equations describes a
periodic oscillation of x about an average value of 0 with an amplitude A (i.e., a maximum
of xmax = +A and a minimum of xmin = −A), angular frequency ω, period T = 2π/ω, and
frequency f ≡ 1/T = ω/2π. The argument of the sine function, (ωt + φ0 ), is called the
phase angle; φ0 by itself is called the phase constant and reflects initial conditions specific to a
particular situation. We also saw in Lecture 1 that writing x = A cos(ωt + Φ 0 ) leads to exactly
the same differential equation linking ẍ and x, and to oscillations with the same amplitude,
period, and so on. The only difference between x(t) written as a cosine and x(t) written as a
sine, is in the value of the phase constant—namely, Φ0 = φ0 − π/2.

Differentiating equation (1) once with respect to time yields

dx
ẋ ≡ = ωA cos(ωt + φ0 ) , (3)
dt
and differentiating once more (or, equivalently, putting eq. [1] into the right-hand side of eq. [2])
gives
d2 x
ẍ ≡ 2 = −ω 2 A sin(ωt + φ0 ) . (4)
dt

Table 1. x(t) = A sin(ωt + φ0 ) and its first two time-derivatives

phase angle, ωt + φ0 (radians)


0 π/2 π 3π/2 2π

x 0 +A 0 −A 0
ẋ +ωA 0 −ωA 0 +ωA
ẍ 0 2
−ω A 0 2
+ω A 0
–2–

We can use these results to compare the values of x, ẋ, and ẍ at a few well-chosen phase
angles, as in Table 1. This shows directly that the points at which x and ẋ achieve their
respective maxima are separated by π/2 radians (or 90◦ ) in phase (xmax = +A is attained at
ωt + φ0 = π/2, while ẋmax = +ωA is achieved when ωt + φ0 = 0). Likewise, the points at
which ẋ and the second derivative ẍ are maximized are also separated by π/2 rad in phase
(ωt + φ0 = 2π for ẋ maximized, versus ωt + φ0 = 3π/2 for maximum ẍ). Thus, we say that ẋ
is π/2 or 90◦ out of phase with x, and that ẍ is π/2 out of phase with ẋ. Note that ẍ is a full
π radians = 180◦ out of phase with x.

Adding any integer multiple of 2π to the phase angle (ωt + φ0 ) leaves the values of all of x, ẋ,
and ẍ unchanged—simply because of the sine or cosine dependence of these quantities on the
phase. Noting this is another way of seeing that the period T of the oscillation in x must be
T = 2π/ω [since then ω(t + nT ) + φ0 = (ωt + φ0 ) + nωT = (ωt + φ0 ) + 2nπ, for any integer
n]. Therefore, we can also say that the first time-derivative ẋ is one-quarter of a cycle (T /4 in
units of time) out of phase with x; that the second time-derivative ẍ is a further quarter-cycle
out of phase with ẋ; and that ẍ is one-half of a cycle (T /2 in terms of time) out of phase with
x itself. See Figure 15.3 of Benson’s University Physics for a diagram illustrating these phase
differences.

In Lecture 1 we arrived at the results above by considering the motion of a point around a circle
about the origin in the (x, y) plane—that is, by beginning in effect with equation (1) for x as
a function of time. However, as we also stressed earlier, the notation and terminology that we
have introduced, and the basic results we have obtained so far, are much more broadly relevant,
holding in many situations that need not involve circular motion in any way. Indeed, there
are important applications in which a sinusoidally time-varying quantity x is not even related
to spatial motion, but could represent some other physical property of a system (e.g., density
or electric-field strength, to give two examples of possible non-mechanical oscillations). We
shall continue for the moment to focus on mechanical oscillations, in which the variable x does
represent a spatial displacement or position (and thus ẋ is a velocity, and ẍ an acceleration),
but we no longer tie our discussion to circular motion per se.

2. Simple Harmonic Motion Defined

In many “real-world” situations we actually have direct information (or hypotheses!) about
the net force F acting on an object or a physical system, rather than explicit knowledge of
position versus time. We therefore frequently begin with an equation for the acceleration
ẍ = d2 x/dt2 = F/m as a function of position x and/or time t, and we need to understand the
implications for x(t), ẋ(t) = dx/dt, the energetics of the system, and so on. In this context,
the equation of motion

d2 x
ẍ = −ω 2 x or + ω2 x = 0 (5)
dt2
is one that arises very frequently, as either an exact or an approximate description of a broad
range of physical systems.
–3–

Any quantity governed by equation (5) is said to undergo simple harmonic oscillation,
or SHO. (For mechanical oscillations specifically, when x is a position, this is also called
simple harmonic motion, or SHM.) The reason for “oscillation” in this name is clear, as
we have already seen that the solution to equation (5) is x(t) = A sin(ωt + φ 0 ) for constants
A and φ0 . The term harmonic means that the sinusoidal variation of x over time involves a
single angular frequency, ω. The term simple refers to the fact that the total energy is
constant, in both space and time. As we shall see in Lecture 3, this conservation of energy is
equivalent to the amplitude A of the oscillation in x being constant.

Given ẍ = −ω 2 x as in equation (5), it can be seen that x = 0 is a point of equilibrium


(zero acceleration means no net force at x = 0). For displacements x > 0, the minus sign in
equation (5) means that ẍ < 0, i.e., the acceleration is such that it opposes further movement
in x away from equilibrium. Conversely, when x < 0, the acceleration ẍ is > 0, which tends
again to bring x back towards the equilibrium position. Thus, simple harmonic motion follows
from the influence of a restoring force, implying that the equilibrium at x = 0 is a stable
equilibrium. But more than this, the restoring force (and thus acceleration) must depend
linearly on the displacement from equilibrium, precisely as it does in equation (5). [If we
instead had, for example, ẍ = −ω 2 x2 instead of −ω 2 x, then x = 0 would still be a point of stable
equilibrium, but the motion in x would not be simple harmonic, since x(t) = A sin(ωt + φ0 )
would no longer satisfy the starting equation for ẍ.]

ASIDE:

We have noted before that the (constant) amplitude A and the phase constant φ 0 in the full
solution for x(t) in simple harmonic motion do not appear in the equation of motion. This
is typical of the solutions to second-order, ordinary differential equations (ODEs), of which
d2 x/dt2 + ω 2 x = 0 is a relatively simple example.

A differential equation is an equation involving one or more derivatives of a quantity


(d2 x/dt2 or ẍ in our case) and some general function of the quantity itself (i.e., the term
ω 2 x in our case, but in general other quantities and/or constants could also be involved). An
ordinary differential equation is one that involves total derivatives, as opposed to partial
derivatives (which we shall encounter later). The term second-order means that the highest
derivative involved in the equation is the second derivative (thus, for example, an equation
such as dx/dt = −ω 2 x would be a first-order ordinary differential equation; but an equation
like d2 x/dt2 + γ(dx/dt) + ω 2 x = 0 is still second-order).

From a mathematical point of view, if we want to derive the solution for x(t) starting from a
second-order ODE such as ẍ = −ω 2 x, we have to integrate twice with respect to time: once to
R R
find ẍ dt = ẋ as a function of time, and then again to find ẋ dt = x as a function of time.
Each step introduces its own constant of integration, and in our case these appear as the two
constants A and φ0 in the general solution x(t) = A sin(ωt + φ0 ).

It is therefore not possible to solve completely for the SHO of a real physical system, given
only the defining equation (5) and some value for the angular frequency ω: the basic form of
equation (1) follows immediately from the differential equation alone, but we always also need
–4–

two additional, independent pieces of information (such as, say, the values of x and ẋ at t = 0)
in order to infer specific values for A and φ0 as well.

To summarize, systems that will undergo simple harmonic oscillation are those in which the
total energy is conserved, a stable equilibrium exists, and the net force is directly proportional
in magnitude and opposite in direction to the displacement from stable equilibrium. In the
remainder of this Lecture we consider two fairly straightforward systems that satisfy all of these
criteria. Other worked examples can be found in Chapter 15 of Benson, University Physics.
In Lecture 3 we will move on to systems for which SHM is an approximate solution to the
equation of motion (the so-called simple pendulum is a template example of this), and we will
also discuss the energetics of SHM.

3. Some Physical Systems Exhibiting SHM

Example 1—Block on a horizontal spring

Figure 15.4 of Benson gives a typical illustration of this simple system. A block of mass m
resting on a frictionless, horizontal surface is attached to the end of a spring, and the other
end of the spring is fixed to, e.g., a vertical wall. All forces and motion in the system are in
the x direction only, and it is assumed that there is no dissipation of energy (i.e., no internal
“friction”) in the spring as it is stretched or compressed.

Call the position of the block x = 0 when the spring is at its equilibrium length. Suppose
the block is pulled away from the wall, stretching the spring, until the block is some distance
x = +A away from the equilibrium position. Hold the block still for a moment and then
release it, so that it moves back towards equilibrium under the influence of restoring spring
force. Hooke’s law for springs says that the restoring force is just proportional to the change
in length from equilibrium, i.e.,
Fspring = −kx
where k is called the spring constant. The force on the block is F = ma = md2 x/dt2 by
Newton’s second law, so with F = Fspring as just given,

k
ma = Fspring =⇒ mẍ = −kx =⇒ ẍ = − x.
m
This has precisely the form of equation (5), ẍ = −ω 2 x, required for the block to undergo SHM
about an equilibrium at x = 0. We need only make the association
p
ω 2 ≡ k/m =⇒ ω = k/m

to conclude immediately that x depends on time as


Ãr !
k
x = A sin(ωt + φ0 ) = A sin t + φ0 .
m
–5–

The period of oscillation for the block also follows immediately, from the angular frequency ω:
p
T = 2π/ω = 2π m/k .

Note that the spring constant k has SI units of N m−1 , which is the same as kg s−2 , so the
units of T are seconds—as required—if the mass m is in kg.

Note also that T is independent of both the amplitude and the phase constant of the x motion.
This is an example of a generic feature of simple harmonic oscillation: the period depends only
on the angular frequency ω, which is completely independent of A and φ0 .

For completeness, the frequency f of the block-spring system is


p
f = 1/T = ω/2π = (1/2π) k/m

which has units of s−1 or Hz. (One Hertz is equal to one cycle per second and thus is an
appropriate alternate unit for f but not for the angular frequency ω, which is in radians per
second.)

In the description of this set-up, we said that the block is brought to its maximum of x max = +A
at t = 0, at which point its instantaneous speed is 0. This defines both the amplitude of the
oscillation (A) and the phase constant φ0 in this example: if x = A at t = 0, then A sin(φ0 ) = A,
so φ0 = π/2 radians. Thus, in this specific case,
Ãr !
k
x(t) = A sin(ωt + π/2) = A cos(ωt) = A cos t ,
m

and the velocity and acceleration of the block follow as


r Ãr !
k k
ẋ = −ωA sin(ωt) = − A sin t
m m

and Ãr !
k k
ẍ = −ω 2 A cos(ωt) = − A cos t .
m m

Example 2—Block hanging from a vertical spring

The left side of Figure 1 shows a spring (with spring constant k) hanging from a fixed point.
It is supposed to be in equilibrium, so its length as illustrated is its natural length, and a y
coordinate has been defined so that y = 0 is the position of the free end of the spring. If
this spring were positioned horizontally and a block of mass m were attached to one end, as
in Example 1 just above, then the angular frequency of simple harmonic motion would be
p
ω = k/m.

Now imagine hanging a block of mass m from the free end of the vertical spring. Either pulling
down or pushing up on the block and then releasing it will produce some up-and-down motion
about an equilibrium position, y = yeq (which we will determine). To characterize this motion,
we identify all forces acting on the system and apply Newton’s second law.
–6–

Fig. 1.— A vertical spring with no mass attached (left; in equilibrium, with the free end defining
y = 0) and with a block of mass m hanging from it (right; not necessarily in equibrium).

At some point (not necessarily the equilibrium point), the vertical block+spring system will
look like the right-hand side of Figure 1, in which the spring is stretched relative to its natural
length and the centre of mass of the attached block is located below the original equilibrium,
i.e., at some negative y < 0.

There are two forces acting on the block. One is the force of gravity,

Fgrav = −mg

with g = 9.81 m s−2 the usual acceleration due to gravity. This force is always directed
downards, as the minus sign in Fgrav = −mg indicates. The other force on the block is the
spring force, which is given by Hooke’s law as

Fspring = −ky .

This force may be directed upwards (when the spring is stretched, so that the block is at y < 0
as in Fig. 1: then Fspring = −ky is > 0), but at other times it may be directed downwards
(when the spring is compressed, so that the block is at y > 0: then Fspring = −ky is < 0). The
minus sign in Fspring = −ky takes care of either possibility.

With minus signs already in place to determine the correct direction of both forces on the
block at any time, the net force is simply the sum

Fnet = Fgrav + Fspring = −mg − ky .

Then, by Newton’s second law,

d2 y
−mg − ky = Fnet = m ≡ mÿ ,
dt2
–7–

so that
k
mÿ = −ky − mg =⇒ ÿ = − y−g .
m

This final expression is close to the form of the defining equation for SHM, ẍ = −ω 2 x, but
there is also an extra term, −g, on the right-hand side. However, because this extra term
is just a constant, the motion of the block ultimately is still SHM. The reason is that the
application of a constant force −mg simply moves the equilibrium position of the block (or
changes the equilibrium length of the spring) to something below y = 0, but otherwise does not
change the basic physics of the situation beyond what it would be for the same block+spring
on a frictionless horizontal surface.

To see this mathematically, we first ask where the equilibrium of the hanging block+spring is.
The equilibrium point is that for which an upwards spring force exactly balances the downwards
gravitational force, giving Fnet = 0. The acceleration is then also 0, so
k mg
ÿ = 0 at equilibrium =⇒ − yeq − g = 0 =⇒ yeq = − .
m k
Notice that yeq < 0, so that—as expected—in the new equilibrium the spring is stretched
beyond its natural length.

Given this expression for yeq , we can write our equation above for the y-acceleration as
k k mk k mg k k ³ mg ´ k
ÿ = − y −g = − y −g × = − y− = − y+ = − (y − yeq ) .
m m k m m k m m k m
This then suggests that we define a new coordinate,

Y ≡ y − yeq ,

which directly measures the position of the block from the equilibrium yeq . Then because yeq
is just a constant, the acceleration in Ÿ is equal to the acceleration in the original ÿ:
d2 Y d2 y d2 yeq d2 y k k
Ÿ = 2
= 2
− 2
= 2
− 0 = ÿ = − (y − yeq ) = − Y .
dt dt dt dt m m
That is, we have shown that the system obeys the equation of motion
k
Ÿ = − Y ,
m
where Y is the vertical distance of the block from equilibrium. This equation of motion is now
exactly of the form ẍ = −ω 2 x, and thus the motion is simple harmonic oscillation about
p
an equilibrium Y = 0 (or y = yeq = −mg/k) with an angular frequency ω = k/m, a period
p p
T = 2π/ω = 2π m/k, and a frequency f = ω/2π = (1/2π) k/m. Again, all of this is the
same as for the block+spring on a horizontal surface. The only difference is that, in
the hanging case, in equilibrium the spring is stretched beyond its natural length.

Because of the form of the differential equation of motion, we know immediately that the
displacement as a function of time must be Y (t) = A sin (ω t + φ0 ); that hence that Ẏ (t) =
p
ωA cos (ω t + φ0 ) and Ÿ (t) = −ω 2 A sin (ω t + φ0 ). The angular frequency ω = k/m is set
directly by the block and the spring, while the constants A and φ0 depend on the exact
configuration of a given block+spring at the time we choose to call t = 0 (initial conditions).
–8–

The “trick” of re-defining a variable Y = y − yeq , so that equilibrium occurs at Y = 0 and the
equation of motion could be cast in our “standard form” (equation [5]) for SHM, is a common
procedure, and you should become familiar with it. We do the same thing in Example 5 below
(a block floating in a fluid).

Example 3—The simple pendulum: approximate SHM

Figure 2 illustrates an idealized pendulum: a bob of mass m at the end of a massless string of
length L, which hangs from a fixed point at its other end. When such a pendulum is allowed
to swing freely (with no air resistance and no friction or dissipation of energy at the fixed
point) through relatively small angles, its motion is approximately simple harmonic.
The derivation of this here expands slightly upon the derivation in Chapter 15.4 of Benson.

Because the bob on the pendulum is constrained to move along a circle of radius L, the natural
coordinate to use in the problem is the arc length s = Lθ, where θ is the angle that the

Fig. 2.— Simple pendulum: defining coordinates and illustrating decomposition of the gravi-
tational force Fy on the bob m to obtain the component Fs along the circular arc s.
–9–

string of the pendulum makes to the vertical at any instant. If the angle θ is measured
in radians, then s = Lθ is the exact distance of the bob from the bottom of its swing,
as measured along the circle of radius L. The bottom of the swing is the obvious point of
equilibrium in this problem, and thus we take s = 0 (and θ = 0) there. Both s and θ are
functions of time, while L is constant.

The motion of the bob is once again analysed by identifying all forces acting on it, and then
applying Newton’s second law. One force acting on the bob is the tension in the string, which
always points directly along the string, away from the bob and towards the fixed point of
attachment at the top. The other force is the force of gravity, which is always F grav = −mg (the
minus sign is to account for the fact that the gravitational force is always straight downwards,
whereas by default y is defined to be positive upwards).

By construction, the bob cannot move at all in the (radial) direction along the string. Thus,
we decompose the gravitational force into its components FL parallel to the string and Fs
perpendicular to this (so tangent to the circular arc s). From the diagram in Figure 2, the
magnitude of the force along the string is FL = mg cos θ and its direction is outwards, away
from the fixed point at the top of the pendulum. This force must be cancelled by the tension
in the string, however, so that the net force in the direction along the string is always exactly
0 (if it were not, the string would stretch or go slack). The only unbalanced force is then the
component of gravity perpendicular to the string, Fs = −mg sin θ (from Figure 2 again). The
minus sign in Fs accounts for the fact that, whenever the bob is away from its equilibrium at
s = θ = 0, gravity always acts to pull the bob back towards the equilibrium.

Thus, this is really a one-dimensional problem, with all motion of the bob being along the
circular arc s and described by Newton’s second law as
d2 s
Fs = m ≡ ms̈ = −mg sin θ =⇒ s̈ = −g sin θ .
dt2
Although this last equation appears to contain two variables, s and θ, we know that s = Lθ
by definition, so that only one of the two is independent. We can thus write the equation of
motion entirely in terms of θ, the angular displacement from equilibrium:
g
s ≡ Lθ =⇒ s̈ = L θ̈ =⇒ L θ̈ = −g sin θ =⇒ θ̈ = − sin θ .
L
The final equation is exact, holding for any value of the angle θ. However, when θ is small
enough, it happens that sin θ ' θ to a good approximation, provided that θ is measured in
radians. (This approximation is good to better than ' 5% for θ . 30◦ , i.e., θ . π/6 rad. . . try
it on a calculator.) In this small-angle limit, we can therefore write
g g
θ̈ = − sin θ ' − θ ,
L L
which shows that the equation of motion is approximately of the basic form θ̈ = −ω 2 θ, with
ω 2 = g/L, and thus the motion of the pendulum is approximately simple harmonic, with
the angular displacement θ oscillating about the equilibrium θ = 0.

The angular frequency of the simple pendulum in the small-angle limit is given directly by the
form of the equation of motion, and the period and frequency of the oscillation follow from ω:
r s r
g 2π L 1 ω 1 g
ω = ; T = = 2π ; f= = = .
L ω g T 2π 2π L
– 10 –

In the small-angle limit, the angular displacement, angular velocity, and angular ac-
celeration of the pendulum are given as functions of time approximately by

θ(t) = θmax sin(ω t + φ0 )


θ̇(t) = ωθmax cos(ω t + φ0 )
θ̈(t) = −ω 2 θmax sin(ω t + φ0 ) = −ω 2 θ(t)
p
where ω = g/L as just discussed; θmax is the angular amplitude, which should be less
than about 30◦ (π/6 rad) if the approximation of SHM is to describe the pendulum with
reasonable accuracy; and φ0 is the phase constant. Both θmax and φ0 must be inferred from
the initial conditions of any particular simple pendulum.

Finally, the linear displacement, velocity, and acceleration of the bob as functions of
time follow simply from the angular quantities: s(t) = L θ(t); ṡ(t) = L θ̇(t); and s̈(t) = L θ̈(t).

Example 4—Other pendulums

Read Section 15.4 of Benson to learn about torsional pendulums and physical pendulums.
Understand the derivation of their equations of motion, and appreciate the difference between
exact SHM and approximate SHM for different kinds of pendulum.

Example 5—Block floating in a fluid

Consider the situation illustrated in Figure 3, in which a block of some density ρb and mass mb
floats in a fluid of density ρf . Let the height of the block be called h (marked in the figure),
and the cross-sectional area (i.e., the area of the top of the block going into/out of the page
and not shown in the figure) be A. Then the mass of the block is

mb = (density) × (volume) = ρb Ah .

(Note that A in this example is an area, not an amplitude of oscillation.)

Let the coordinate y measure the vertical distance from the surface of the fluid, which is shown
as the dashed line at y = 0 in Figure 3. If the bottom of the block is some distance y < 0 below
the surface, then the forces acting on the block are the (downwards) gravitational force,

Fgrav = −mb g = −ρb Ahg

and the (upwards) buoyancy force due to the fluid trying to support the clock. The buoyancy
force is equal to the weight of an amount of fluid with the volume of the submerged part of
the block (since if the block were removed and the volume between the block bottom and the
fluid surface were filled back in with fluid, the fluid would be in equilibrium, with buoyancy
balancing gravity exactly). Thus,

Fbuoy = (density of fluid) × (volume of submerged part of block) × g


= + ρf × A|y| × g
= − ρf Agy .
– 11 –

Fig. 3.— Block floating in a fluid. The mass of the block is given in terms of its density, height,
and cross-sectional area A (into the page; not shown) as mb = ρb Ah.

The minus sign in the last line comes from the fact that y < 0 when the bottom of the block
is below the surface of the fluid; that is, |y| = −y.

The net force on the block is just the sum of Fgrav and Fbuoy :

Ftotal = mb × d2 y/dt2 = (ρb Ah) ÿ = −ρb Ahg − ρf Agy ,

which can be rearranged to isolate the acceleration ÿ on the left-hand side:


ρf g
ÿ = −g − y. (6)
ρb h
Notice that the cross-sectional area A of the block has cancelled out of the equation of motion
for y. What is left looks nearly like an equation leading to simple harmonic oscillation in y (i.e.,
one with the basic form ÿ = −ω 2 y). The only difference is the addition of another, constant
term of −g to the y-dependent term in the acceleration. However, this only reflects the fact
that the equilibrium point of this system is not at y = 0 (as our derivations to this point have
assumed, for simplicity), but must be at some y < 0—that is, in equilibrium at least part of
the block must be submerged.

We can nevertheless infer the position yeq of the bottom of the block when it is in equilibrium,
and then look at the equation of motion for the block around this equilibrium. Thus, first set
ÿ = 0 in equation (6) and solve for the equilibrium position:
ρb
ÿ = 0 =⇒ yeq = − h. (7)
ρf
– 12 –

Notice that yeq < 0, as it must be (the bottom of the block must be below the surface of the
fluid). Also, if the block and the fluid have the same density, ρb = ρf , then yeq = −h—in
which case the top of the block is at y = 0, i.e., exactly at the surface of the fluid. If ρ b > ρh
then yeq < −h so even the top of the block must be submerged, and a “floating” equilibrium
is not possible. Only when ρb < ρf , so that 0 > yeq > −h, can the block be only partially
submerged in equilibrium.

In any event, given equation (7) defining yeq , we can re-write equation (6) as
ρf g
ÿ = −g − y
ρ h
µb ¶
ρf g ρb
= − y+ h
ρb h ρf
ρf g
= − (y − yeq ) . (8)
ρb h
Finally, we define a new position variable

Y ≡ y − yeq

to measure the distance of the floating block from its equilibrium at y = yeq . By definition,
then, Y = 0 in equilibrium. Moreover, any time derivative of the new Y is exactly the same
as the corresponding time derivative of the old y, because yeq is strictly constant and all
its derivatives are 0. In particular, the second derivative Ÿ is equal to the original second
derivative ÿ. Therefore, equation (8) can be re-written again, as
µ ¶
ρf g
Ÿ = − Y . (9)
ρb h

This is now of exactly the general form (5) for simple harmonic motion. Thus, from consider-
ation of forces (and the associated equilibrium) alone, we have found that a block of uniform
density ρb and height h floating in a fluid of density ρf > ρb will, if displaced from equilibrium,
undergo simple harmonic motion about an equilibrium position of yeq = −(ρb h/ρf ) for the
block bottom (i.e., an equilibrium Y = 0), with an angular frequency

ρf g g
r r
ω≡ = . (10)
ρb h |yeq |

The associated period of oscillation is the usual


s s
2π ρb h |yeq |
T = = 2π = 2π , (11)
ω ρf g g

which is again independent of either the amplitude or the phase constant of the oscillation
(both of which we left unspecified in this example).

It is worth noting the similarity of equations (10) and (11) to the analogous results for ω and
T of a simple pendulum (above): they are formally the same, with |yeq | taking the place of
L. Of course, our results for the floating block are exact, while the oscillation of the simple
pendulum is only approximate SHM in the limit of small angular amplitude.
Lecture 3 Sample Problems on SHM

The first section of these notes (pp. 1–8) consists of 8 worked problems dealing with simple
harmonic motion. These are slightly modified versions of some exercises from the end of
Chapter 15 in the Benson text; you should attempt to solve several other problems from the
text on your own (answers to odd-numbered problems are in the back of the book).

The second section (pp. 8–10) develops a more sophisticated application, showing that stars in
an idealised, spherical and uniform star cluster undergo simple harmonic motion and estimating
the period of their oscillation about the centre.

1. Sample Problems on Simple Harmonic Motion

Problem 1: A particle moves in simple harmonic motion about x = 0 with period


0.4 s. At t = 0, it has its maximum positive acceleration, 28 m s−2.
(a) Find the amplitude and the phase constant of the oscillation. (b)
Write the displacement as a function of time.

Because the particle is in SHM, we know immediately that

x(t) = A sin(ωt + φ0 )
ẋ(t) = ωA cos(ωt + φ0 )
ẍ(t) = −ω 2 A sin(ωt + φ0 )

(a) First, we are told that the period is T = 0.4 s, so from the defining relation between
T and ω we know that
2π 2π
T = = 0.4 s =⇒ ω= = 5π s−1 .
ω 0.4 s
Then, we are told that at t = 0, ẍ is > 0 and a maximum. The maximum possible value
for ẍ is ẍmax = +ω 2 A, which occurs whenever sin(ωt + φ0 ) = −1. Since this happens when
t = 0, we therefore know that sin(φ0 ) = −1, which means

φ0 = 3π/2 radians .

(This is also φ0 = 270◦ , or φ0 = −90◦ , or φ0 = −π/2 rad.)


Finally, we are given that the maximum value of ẍ is 28 m s−2 , so

ẍmax 28 m s−2 28
ẍmax = ω 2 A = 28 m s−2 =⇒ A = 2
= 2
= m ' 0.113 m .
ω (5π s )
−1 25π 2

(b) Knowing all of ω, φ0 , and A from part (a), and the general form of x(t) for any SHM,
–2–

we have

x(t) = A sin(ωt + φ0 )
µ ¶
28 3π
= sin 5π t + m
25π 2 2
28
= − cos (5π t) m
25π 2

Problem 2: A point in the middle of a guitar string oscillates with frequency 440
Hz and amplitude 0.8 mm. (This frequency corresponds to the note
A above middle C.) (a) What is the maximum speed of the point? (b)
What is the maximum acceleration of the point?

Before calculating any numbers, recall that the displacement, velocity, and acceleration of
any simple harmonic oscillator are

y(t) = A sin(ωt + φ0 )
ẏ(t) = ωA cos(ωt + φ0 )
ÿ(t) = −ω 2 A sin(ωt + φ0 )

(a) In this particular case, we are given f = 1/T = 440 Hz = 440 s−1 (we’re told the
“frequency,” not the “angular frequency”). So ω = 2π/T = 2πf = 880π s−1 . Then,
from the general expressions immediately above, we know that the maximum speed is
|ẏ|max = ωA [achieved at times such that cos(ωt + φ0 ) = ±1]. In this case, the amplitude
A = 0.8 mm, so the maximum speed is

|ẏ|max = ωA = (880π s−1 ) × (0.0008 m) = 2.21 m s−1 .

(b) From ÿ = −ω 2 A sin(ωt + φ0 ), the maximum acceleration (absolute value) is |ÿ|max =


ω 2 A in general. Here,

|ÿ|max = ω 2 A = (880π s−1 )2 × (0.0008 m) = 6114 m s−2 .

Problem 3: A 10-gram mass is attached to a horizontal spring with spring constant


k = 1.21 N m−1. It is pulled 5 cm out from its equilibrium position
and released at t = 0. Write x(t).

We know that

x(t) = A sin(ωt + φ0 )
ẋ(t) = ωA cos(ωt + φ0 )
ẍ(t) = −ω 2 A sin(ωt + φ0 )

for any system undergoing SHM. Because this specific system is a block on a spring, we
also know that r s
k 1.21 N m−1
ω = = = 11 s−1 .
m 0.010 kg
–3–

We thus need only to find A and φ0 in order to determine x(t) completely. These constants
follow from the information we have on the system at t = 0.

First, the position at t = 0 is x(0) = A sin φ0 , which we are told is equal to 5 cm. We
are also told that the block is “released” at t = 0, meaning that ẋ(0) = 0. Thus, we
infer that cos φ0 = 0 which means that either φ0 = π/2 (equivalently, 90◦ ) or φ0 = 3π/2
(equivalently, φ0 = 270◦ or −90◦ .)

If we had that φ0 = 3π/2 (or 270◦ , or −90◦ , or −π/2 rad), then we would also have that
sin φ0 = −1, and x(0) would have to be negative. But we are told the the block is “pulled
out” from equilibrium at t = 0, so that x(0) must be positive. We therefore must have

φ0 = π/2

(or φ0 = +90◦ ) in order to ensure sin φ0 = +1. Then we immediately also have x(0) =
A sin φ0 = A, so the amplitude of oscillation is

A = 5 cm = 0.05 m .

Putting these things together:

x(t) = 0.05 sin (11 t + π/2) m = 0.05 cos (11 t) m .

Problem 4: A block of mass 60 g is attached to a spring with k = 24 N m−2. The


spring is extended, held at rest and then released at t = 0. Later, at
t = 0.05 s, the velocity is ẋ = −0.69 m s−1. What is the amplitude of
the oscillation?

Starting again from the general expressions,

x(t) = A sin(ωt + φ0 )
ẋ(t) = ωA cos(ωt + φ0 )
ẍ(t) = −ω 2 A sin(ωt + φ0 )

This particular system is released from rest at t = 0, so the fact that ẋ(t = 0) = ωA cos(0+
φ0 ) in general means that cos φ0 = 0 here, and thus either φ0 = π/2 rad (90◦ ) or φ0 =
3π/2 rad (270◦ or −90◦ ). In the first case, sin φ0 = +1, so x(t = 0) = A sin φ0 = +A. In
the second case, sin φ0 = +1 and x(t = 0) = A sin φ0 = −A. The first case is the correct
one in this example, since we are told that the spring is “extended” (not compressed) at
t = 0, meaning x(0) must be positive. That is,

φ0 = π/2 rad (or + 90◦ ) .

Given this, it is more convenient in this case to write

x = A sin(ωt + π/2) = A cos(ωt) ,

and then the velocity becomes

ẋ = −ωA sin(ωt) .
–4–

Now, at t = 0.05, ẋ = −0.69 m s−1 , so from the formula immediately above we can write

−0.69 m s−1 = −ωA sin(0.05ω) .


p
We are not given ω explicitly, but we know that ω = k/m for a spring, and we have the
values of k and m. Therefore,
s
24 N m−1
ω = = 20 s−1
0.060 kg

and so at t = 0.05,

−0.69 m s−1 = −20 s−1 × A sin(0.05 × 20) .

Solving for the amplitude,

−0.69 m s−1
A = = 0.0410 m = 4.10 cm .
−20 s−1 × sin(1)

Problem 5: A simple pendulum, with length 0.4 m and a bob of mass 50 g, is


released when it makes an angle of 20◦ to the vertical. Find (a) the
period of its swing; (b) the angular displacement as a function of time;
(c) the speed of the bob at the lowest point; (d) the total energy.
p
(a) The period of any simple pendulum is T = 2π L/g, so in this case,
p p
T = 2π L/g = 2π (0.4 m)/(9.81 m s−2 ) = 1.27 s .

(b) The angular displacement of a simple pendulum and its first two time derivatives are

θ(t) = θmax sin(ωt + φ0 )


θ̇(t) = ωθmax cos(ωt + φ0 )
θ̈(t) = −ω 2 θmax sin(ωt + φ0 )
p
where θmax is the angular amplitude, φ0 is the phase constant, and ω = g/L if the length
of the pendulum is L.

In this case, the pendulum is “released” at t = 0, so the angular velocity then is θ̇(0) = 0.
This requires cos φ0 = 0, and so either φ0 = π/2 or φ0 = 3π/2. At the same time,
the actual angle at t = 0 is θ(0) = θmax sin φ0 in general, and we are told that this is
θ(0) = 20◦ —implying that it is positive. It must therefore be that sin φ0 = +1, which
disallows φ0 = 3π/2, leaving
φ0 = π/2 rad
and then
θmax = θ(0)/ sin φ0 = 20◦ /1 = (π/9) radians .
In addition, the angular frequency for this pendulum is
p p
ω = g/L = (9.81 m s−2 )/(0.4 m) = 4.952 s−1
–5–

so, finally,

θ(t) = θmax sin(ωt + φ0 ) = (π/9) sin(4.952 t + π/2) rad = (π/9) cos(4.952 t) rad .

(c) Refer to the general equations for θ(t) and its derivatives, written out in part (b). At
the lowest point of the swing, θ = 0, which implies that sin(ω t + φ0 ) = 0 and thus that
cos(ω t + φ0 ) = ±1. This then means that the angular velocity at this point is either its
minimum, θ̇min = −ω θmax , or its maximum, θ̇max = +ω θmax . Either way, the angular
speed |θ̇| is maximised at θ = 0 (equilibrium):

|θ̇|max = ω θmax ,

which in this particular case is

|θ̇|max = (4.952 s−1 ) × (π/9 rad) = 1.729 s−1 .

This corresponds to a maximum linear speed of

|ṡ|max = L |θ̇|max = 0.4 m × 1.729 s−1 = 0.691 m s−1 ,

which of course also occurs at the bottom of the swing.

(d) The total energy follows from knowing the speed at the bottom of the swing, since
there is no potential energy at that point (we generally define potential-energy scales so
that U = 0 at equilibrium). That is,
1 1
Etot = m|ṡ|2max = × 0.050 kg × (0.691 m s−1 )2 = 0.0120 J .
2 2

Problem 6: A simple pendulum is 0.8 m long and has a bob weighing 20 grams.
It is released from an angle of 30◦ to the vertical. Find (a) its period;
(b) the angular displacement as a function of time; (c) the speed of
the bob at θ = 15◦.
p
(a) T = 2π L/g = 1.79 s .

(b) By the same type of argument as in Problem 5(b) above, the angular frequency is
p
ω = g/L = 3.50 s−1 ; the phase constant is φ0 = π/2, and the angular amplitude is
θmax = 30◦ = π/6 radians. Thus,

θ(t) = 30 sin(3.50 t+π/2) degrees = 30 cos(3.50 t) degrees = (π/6) cos(3.50 t) radians .

(c) Given the expression for θ(t) in part (b), the angular velocity is (using the chain rule
of differentiation!)

θ̇(t) = dθ/dt = −3.50 × (π/6) sin(3.50 t) s−1 .

To find the value of this when θ = 15◦ (which is π/12 radians), it is not necessary to
solve explicitly for the times (plural!) at which θ = 15◦ —although of course this could
–6–

be done. Rather, we need only to notice that θ = 15◦ = π/12 rad is exactly one-half
the full angular amplitude (θmax = 30◦ = π/6 rad). Thus, a displacement of 15◦ is
achieved at any time such that cos(3.50 t) = 1/2. But then, at this displacement we have
p p √
sin(3.50 t) = 1 − cos2 (3.50 t) = 1 − (1/2)2 = ± 3/2 (because sin2 + cos2 = 1 for any
angle). Therefore,

θ̇(t) = −3.50 × (π/6) sin(3.50 t) s−1 =⇒



|θ̇| = 3.50×(π/6)×( 3/2) = 1.587 s−1 when θ = 15◦ = π/12 rad .
This is the angular speed. The corresponding linear speed is |ṡ| = L |θ̇| in general, so at
this particular position,

|ṡ| = 0.8 m × 1.587 s−1 = 1.27 m s−1 .

Problem 7: A simple pendulum has angular amplitude θmax = 20◦ and a period
of T = 2 seconds. How long does it take to swing from θ = −10◦ to
θ = +10◦?

The angular displacement of a simple pendulum has the general form

θ = θmax sin(ωt + φ0 ) .

Suppose θ = −10◦ at time t = t1 and θ = +10◦ at time t = t2 ; then we want to find the
time difference (t2 − t1 ). Given that θmax = 20◦ , at t1 we have
1 π
−10◦ = 20◦ sin(ωt1 + φ0 ) =⇒ sin(ωt1 + φ0 ) = − =⇒ (ωt1 + φ0 ) = − [rad] .
2 6
Similarly, at t = t2 ,
1 π
+10◦ = 20◦ sin(ωt2 + φ0 ) =⇒ sin(ωt2 + φ0 ) = + =⇒ (ωt2 + φ0 ) = + [rad] .
2 6
Subtracting these,
π ³ π´ π
(ωt2 + φ0 ) − (ωt1 + φ0 ) = ωt2 − ωt1 = ω(t2 − t1 ) = + − − = .
6 6 3
Now, we aren’t told the angular frequency directly, but we are given the period and we
know that ω = 2π/T , so
1π T π T 2s 1
t2 − t 1 = = = = = sec .
ω3 2π 3 6 6 3

Problem 8: A vertical spring is extended beyond its natural (zero-load) length by


0.16 m when a block of mass 0.5 kg is hung from it. The block is then
pulled down a further 0.08 m and then released (from rest).
(a) Write an equation for the displacement x from equilibrium, as a
function of time.
–7–

(b) Find the speed and acceleration of the block when the spring is
extended by a total of 0.1 m beyond its natural length.

(a) From Lecture 2, we know that the equilibrium stretch of a hanging spring, beyond its
natural length, is
mg
|yeq | =
k
when a block of mass m is hung from it. We also proved in Lecture 2 that the angular
frequency ω of simple harmonic oscillation of such a system is the same as for the same
spring+block on a horizontal surface. That is,
r
k g
r
ω = = .
m |yeq |

In this example, we are told that |yeq | = 0.16 m; and we know that the acceleration due
to gravity is g = 9.81 m s−2 , so
r
g 9.81 m s−1
r
ω = = = 7.830 s−1 .
|yeq | 0.16 m

Now, let the displacement of the hanging block from its equilibrium position be denoted
x ≡ (y − yeq ), with y defined so that y = 0 marks the end of the spring when it is at its
natural length. From Lecture 2 again, the equation of motion in terms of this x is

x ≡ (y − yeq ) =⇒ ẍ = −ω 2 x ,

so the full dependence of x on t is the usual

x(t) = A sin(ωt + φ0 ) ,

with the associated velocity and acceleration,

ẋ(t) = ωA cos(ωt + φ0 )
ẍ(t) = −ω 2 A sin(ωt + φ0 )

We are given that x = −0.08 m at t = 0 (negative because the block is pulled below its
equilibrium initially), and that the velocity is ẋ = 0 at t = 0. By logic similar to the other
examples above, this information implies

A = 0.08 m and φ0 = 3π/2 (or 270◦ , or − 90◦ )

for the amplitude and the phase constant of this specific oscillation. Thus,

x(t) = 0.08 sin(7.830 t + 3π/2) m = −0.08 cos(7.830 t) m .

(b) From part (a),


x(t) = −0.08 cos(7.830 t) m ,
so the velocity and acceleration of the block are

ẋ = +7.830 × 0.08 sin(7.830 t) = +0.6264 sin(7.830 t) m s−1


–8–

and
ẍ = +(7.830)2 × 0.08 cos(7.830 t) = +4.905 cos(7.830 t) m s−2
When the spring is stretched by 0.1 m from its natural length, the position of its end is
y = −0.1 m, and the displacement from equilibrium is x = y − yeq = −0.1 − (−0.16) =
+0.06 m. At this point,
0.06 3
x = −0.08 cos(7.830 t) = +0.06 =⇒ cos(7.830 t) = − =−
0.08 4
r
p 7
=⇒ sin(7.830 t) = 1 − cos2 (7.830 t) = ± .
16
Hence, the speed there is
r
7
|ẋ| = 0.6264 × | sin(7.830 t)| = 0.6264 × = 0.414 m s−1
16
and the acceleration is
µ ¶
3
ẍ = 4.905 cos(7.830 t) = 4.905 × − = −3.679 m s−2 .
4

2. Motion in a Star Cluster

Within galaxies there are many clusters of stars. Any one cluster may contain thousands to
millions of stars, each star being similar in mass to our Sun (on average), and all confined to
a relatively small volume (in astronomical terms!) by the force of their mutual gravitational
attractions. In this section, we use a very simple model for such star clusters, and our knowledge
of basic SHM, to gain some insight into how individual stars move within the clusters.

Suppose a cluster consists of N stars, each with the same mass M? , all packed into a volume
of radius R. We show now that, if the cluster is modeled as a homogeneous (uniform-density)
sphere, then the motion of stars on radial orbits inside the cluster is inferred to be simple
harmonic; and we evaluate the period of the motion for parameters N , M ? , and R typical of
real clusters.

First, the total mass of the cluster is Mtot = N M? , while its total volume is V = (4/3)πR3 .
Thus, the average mass density inside the cluster is
Mtot 3N M?
ρave = =
V 4πR3

Assume that the cluster as a whole is spherical and roughly uniform-density; that is, picture
all the individual member stars as smeared out to give the same total mass density of ρ(r) =
ρave = constant at any position inside the cluster. If a “test mass” m (which could be anything
in principle, but it’s most realistic to think of it as another star) is now located at some radius
r ≤ R, then the net force acting on the test mass comes only from the gravitational attraction
of the mass of the cluster inside the sphere r (the net force exerted by all the cluster mass
outside r is 0). The cluster mass inside the radius of the test star is density × volume:
4 3N M? 4πr3 N M? 3
M (r) = ρave × πr3 = 3
× = r ,
3 4πR 3 R3
–9–

and thus the gravitational force on the test mass is


µ ¶
GmM (r) Gm N M? 3 GN M? m
Fgrav = − 2
=− 2 3
r =− r.
r r R R3

This force acts in the radial (r) direction, i.e., along an imaginary line joining the test mass
and the cluster centre. The minus sign indicates that the force is attractive, always pulling
towards the centre (in the direction of decreasing r).

First consider the case of a test mass on a radial orbit, which is confined to move only along a
straight line passing through the centre of the cluster (no rotation or any other “complicated
motion). On such an orbit, any net force can only be in the radial direction, so analyzing the
motion in this direction alone suffices to describe the entire motion of the test mass.

In this case, Newton’s second law says that

d2 r d2 r GN M? m GN M?
Fnet = m = Fgrav =⇒ m ≡ mr̈ = − r =⇒ r̈ = − r.
dt2 dt2 R3 R3
Notice that m has cancelled out of the problem, so the following results hold for a test particle
of any mass.

The last form of the equation of motion involving the radial acceleration, r̈, is of the generic
form ẍ = −ω 2 x. Thus, in a uniform-density, spherical star cluster, stars on radial orbits execute
simple harmonic motion with a natural angular frequency
r
GN M?
ω=
R3
and hence a period s
2π R3
T = = 2π .
ω GN M?
The oscillation is about r = 0, the centre of the cluster, where r̈ = 0 and which is therefore
the point of (stable) equilibrium.

In general, then, the distance of the star from the centre of such a cluster is given as a function
of time by r(t) = A sin(ωt + φ0 ) with ω as just given and with the amplitude A (the size of
the orbit) and the phase constant φ0 determined from additional information such as initial
conditions. As for all SHM, the angular frequency and the period are independent of any such
amplitude and phase constant for any given star. Again, the results here are also independent
of the mass of the object undergoing the SHM.

Interestingly, if we instead consider stars on circular orbits about the cluster centre, we ulti-
mately find the same orbital period, T . This follows because in this case the motion of a test
mass is described completely by equating the centripetal force associated with a circular orbit
of radius r to the attractive gravitational force from the cluster mass inside r:

mv 2 GN M? m GN M? 2
Fcent = Fgrav =⇒ − = − r =⇒ v2 = r .
r R3 R3
Here v is the orbital speed of any test mass. The direction of motion along a circular orbit is
always perpendicular to the radial direction, and thus v/r is the orbital angular speed. Our
– 10 –

results so far show quickly that this is a constant,


r
v GN M?
= ,
r R3
for a circular orbit in a uniform-density cluster. Thus, the orbital period is simply the change
in angle for one complete orbit (i.e., 2π rad) divided by the constant angular speed:
s
2π R3
T = = 2π .
v/r GN M?

As claimed, this is the same as the period derived above for a radial orbit. In the circular-orbit
case as well, the answer is independent of the radius of the orbit and the mass of the body in
orbit.

Representative values of of N , M? , and R for real star clusters are N = 105 , M? = 1.4×1030 kg
(which is 0.7 times the mass of our own Sun), and R = 10 light-years (one light-year is equal
to the speed of light times the number of seconds in one year: 9.47 × 1015 m). Given these
values (and Newton’s G = 6.67 × 10−11 N m2 kg−2 ), the period of a radial stellar orbit (or a
circular one!) is
s s
R3 (10 × 9.47 × 1015 )3
T = 2π = 2π = 5.992 × 1013 s ,
GN M? (6.67 × 10−11 )(105 )(1.4 × 1030 )

which is also
T = 1.9 × 106 years

Star clusters can live for several billions of years (one billion = 109 ). Thus, stars on radial
orbits in clusters can be expected to cross through the centre many thousands of times during
their lives, while stars on circular orbits make just as many complete loops.
Lecture 4 Energy in Simple Harmonic Motion

1. Energy in Simple Harmonic Motion

We have seen that the equation for simple harmonic oscillation,

d2 x
ẍ ≡ = −ω 2 x , (1)
dt2
has the general solution
x(t) = A sin(ωt + φ0 ) (2)
for constants A (the amplitude of oscillation) and φ0 (the phase constant). This then leads to
dx
≡ ẋ = ωA cos(ωt + φ0 ) (3)
dt
d2 x
and ≡ ẍ = −ω 2 A sin(ωt + φ0 ) = −ω 2 x . (4)
dt2
We now consider the implications of these equations for the energetics of a system undergoing
mechanical oscillations, in which x is a spatial displacement from an equilibrium (at x = 0).

Consider a particle of mass m located at some position x0 relative to equilibrium in a (conser-


vative) force field, where it feels a linear restoring force as implied by equation (1):

Frestore = mẍ0 = −mω 2 x0 .

Now suppose the particle moves by an infinitesimal amount dx0 , from position x0 to position
x0 + dx0 . As it does so, the restoring force does an increment of work on the particle:

dW = Frestore dx0 = −mω 2 x0 dx0 .

Thus, the total work done by the force on the particle as it moves from equilibrium to any
position x is
x x · ¸x0 =x
1 1
Z Z
2
W = dW = − 2 0
mω x dx = − mω 2 x0 + C
0
= − mω 2 x2 .
0 0 2 x0 =0 2

(C is a constant of integration, which we do not need to know.) Now, in general, work done
on a system by a conservative force results in a change of potential energy of the system,

∆U = −W .

The minus sign here arises because if we wanted to move a particle between two positions
against the intrinsic force without acceleration, we would have to apply an external force
of exactly −Frestore to the particle, and the work −W done by that external force corresponds
–2–

to potential energy (see Chapter 8 of Benson, University Physics). In the particular case of
SHM, the formulae above mean that the difference in potential energy between equilibrium
(x = 0) and any other position in the field of a linear restoring force is
1
∆U = U (x) − Ueq = + mω 2 x2 ,
2
where the equilibrium value of the potential energy, Ueq = U (0), is just a constant number.
Re-arranging this, the potential energy at any position x is simply
1
U (x) = mω 2 x2 + Ueq . (5)
2

Because Ueq is just a constant, we are generally free to give it a value that suits our convenience
in any particular problem. In straightforward applications of SHM, it is often taken to be
Ueq = 0. In this case, a standard result is that the potential energy of a particle of mass
m in simple harmonic motion about x = 0 with angular frequency ω is
1
U = mω 2 x2 . (6)
2
Meanwhile, the kinetic energy is the usual
1
K = mẋ2 , (7)
2
since ẋ ≡ dx/dt by definition. Using equations (2) and (3) for the displacement x(t) from
equilibrium and the velocity ẋ(t), we therefore have that the potential and kinetic energies as
functions of time for a particle in SHM are
1
U = 1 2 2
2 mω x = mω 2 A2 sin2 (ω t + φ0 )
2
1
K = 1
2 mẋ
2 = mω 2 A2 cos2 (ω t + φ0 )
2
Clearly, both energies vary in time and neither is conserved on its own. However, the fact that
K varies as cos2 (ω t + φ0 ) while U varies as sin2 (ω t + φ0 ) means in effect that any change in
kinetic energy from one phase (or time) of an oscillation to another will always be balanced by
an exactly opposite change in potential energy—and vice versa (see Figure 15.9 of Benson).
That is, the total mechanical energy, Etot = K + U , is constant:
1 1
Etot = K + U = mẋ2 + mω 2 x2
2 2
1 1
= mω A cos2 (ωt + φ0 ) + mω 2 A2 sin2 (ωt + φ0 )
2 2
2 2
1
mω 2 A2 cos2 (ωt + φ0 ) + sin2 (ωt + φ0 )
£ ¤
=
2
1
=⇒ Etot = mω 2 A2 . (8)
2
In the last line of this derivation, we have used the fact that sin2 θ + cos2 θ = 1 for any angle
θ. The final value for Etot in equation (8) also requires that the amplitude A be
that of a spatial displacement, measured in units of length (it cannot be an angular
amplitude, for example; if it is not a length, then Etot will not even have the dimensions of
–3–

energy). And, again, it also assumes that we have chosen the potential-energy scale
such that U = 0 at the equilibrium x = 0.

Thus, in simple harmonic motion, the total mechanical energy depends on the square of
the angular frequency ω and on the square of the amplitude A of the oscillation. Since
both of these things are constant, energy is conserved in SHM. Conversely, conservation of
energy requires that the oscillation amplitude be constant. Note the correspondence
between equation (8) for Etot and the result that we derived in Lecture 1, for a particle on a
circular “orbit” about the origin in the (x, y) plane.

The quadratic form of the potential energy U as a function of the displacement x in equation
(6) (or, more generally, in equation [5]) is a characteristic of any simple harmonic motion. As
a result of this and the conservation of total energy, the kinetic energy can also be written as
a quadratic function of position:
1 1 1
K = Etot − U = mω 2 A2 − mω 2 x2 = mω 2 A2 − x2 ,
¡ ¢
(9)
2 2 2
which is an inverted parabola. Figure 1 illustrates the dependences of U , K, and E tot on
position in SHM (see also Figure 15.8 of Benson).

Comparing equation (6) for U and equation (8) for Etot shows that the potential energy is
equal to the total energy, U = Etot , at both x = +A and x = −A, i.e., at the endpoints of the
oscillation, when the displacement x is maximized. It must therefore be that K = E tot − U = 0
when x = ±A. (Clearly, this means that the velocity ẋ = 0 when x = ±A—which we already
knew—while we also know that the acceleration is always directed towards x = 0; thus, the
maximum and minimum displacements x = ±A are also referred to as the turning points
of the motion.) By contrast, equation (9) says that K = Etot = (1/2)mω 2 A2 at the point of
equilibrium, x = 0, where the potential energy U = (1/2)mω 2 x2 = 0.

Fig. 1.— Schematic dependence of energy on displacement from equilibrium (x = 0) for


simple harmonic motion with an amplitude of A (−A ≤ x ≤ +A). The potential energy is
U = (1/2)mω 2 x2 (solid red curve), the total energy Etot = (1/2)mω 2 A2 is constant in time and
space (solid, black, horizontal line), and the kinetic energy is K = Etot −U = (1/2)mω 2 (A2 −x2 )
(dashed blue curve). Large black dots mark the positions x where the kinetic and potential
energies are equal.
–4–

In general, then, in SHM the potential energy is a maximum when the kinetic energy
is minimum; and the potential energy is a minimum when the kinetic energy is
maximum. This “competition” between kinetic and potential energies follows from the fact
that U ∝ x2 and K ∝ ẋ2 by definition (equations [6] and [7]), plus the fact that the velocity ẋ
is 0 when the displacement x is at one of the extremes ±A while the displacement is 0 when
the velocity is at one of its extremes ±ωA. Put another way, it follows from the fact that x
and ẋ are out of phase by π/2 rad, or one-quarter of a cycle, in SHM (see the discussion of
this in Lecture 2).

Figure 1 shows explicitly that U = Etot and K = 0 at x = +A and at x = −A, while K = Etot
and U = 0 at x = 0. One complete cycle of oscillation involves the motion of a particle from,
e.g., x = +A to x = −A and then back again, so each of these endpoints is achieved once
during a full cycle; thus, U is maximized and K minimized twice per cycle. Any single
position between −A < x < +A is passed through twice per cycle (once on the way to −A,
and once on the way back from −A). This includes the equilibrium x = 0, and therefore U
is minimized and K maximized twice per cycle. The black dots in Figure 1 mark the
positions x at which the potential and kinetic energies are equal. There are two values of x
for which this occurs, and each of these two positions is passed through twice in an oscillation
cycle; thus, K = U four times per cycle. We can solve for the points in an oscillation at
which this occurs, by setting
1 1
K=U =⇒ mẋ2 = mω 2 x2 =⇒ ẋ2 = ω 2 x2 ,
2 2
which then requires (using equations [2] and [3])

ω 2 A2 cos2 (ωt + φ0 ) = ω 2 A2 sin2 (ωt + φ0 ) for K = U .

The common factor ω 2 A2 on both sides of this equation cancels, leaving


sin2 (ωt + φ0 )
cos2 (ωt + φ0 ) = sin2 (ωt + φ0 ) =⇒ tan2 (ωt + φ0 ) ≡ =1 for K = U .
cos2 (ωt + φ0 )
That is,
tan(ωt + φ0 ) = +1 or tan(ωt + φ0 ) = −1 for K = U ,
which occurs at the phases

ωt + φ0 = π/4, 3π/4, 5π/4, and 7π/4 rad


= 45◦ , 135◦ , 225◦ , and 315◦ for K = U .

Notice that the phase difference between any two consecutive points at which K = U is always
π/2 radians (= 90◦ , or one-quarter of a cycle). Since ω is the (constant) rate of change of the
phase, it takes a time π/2ω = T /4 (the period is T = 2π/ω, as usual) to move from any point
with K = U to the immediate next point with K = U . Finally, given that x = A sin(ωt + φ0 ),
the positions or displacements at which K = U are
A A
x = + √ ' +0.7071 A or x = − √ ' −0.7071 A for K = U .
2 2

The following examples show how some of the ideas above can be applied:
–5–

EXAMPLE At a certain point in the oscillation of a block+spring system the kinetic


energy is 0.1 J and the potential energy is 0.3 J. The amplitude is 20 cm and the period
is 0.8 s. Find the mass of the block and the spring constant.

The total energy is Etot = K + U = 0.1 J + 0.3 J = 0.4 J. This is conserved, and it is equal
to (1/2)mω 2 A2 . We are told that A = 0.20 m and we infer ω from the period: T = 2π/ω
so ω = 2π/T = 2π/(0.8 s) = 2.5π s−1 . Thus, the mass of the block follows from

1 2 Etot 2 × 0.4 J
Etot = mω 2 A2 =⇒ m = = = 0.324 kg .
2 ω 2 A2 (2.5π s−1 )2 (0.20 m)2
p
We also know that ω = k/m for a block on a spring, and so mω 2 = k for this system.
Therefore,
1 1 2 Etot 2 × 0.4 J
Etot = mω 2 A2 = kA2 =⇒ k = = = 20 N m−1 .
2 2 A2 (0.20 m)2

EXAMPLE A 50-gram block is attached to a vertical spring whose stiffness constant is


4 N m−1 . The block is released at the position where the spring is unextended. (a) What
is the extension of the spring when the block is in equilibrium? (b) What is the maximum
extension of the spring? (c) How long does it takes for the block to reach the lowest point?

We use considerations of energy to answer this question. Compare with the alternate,
force-based discussion in Lecture 2 of a block hanging from a vertical spring.

For this problem it is most convenient to define y = 0 as the position of the block when the
spring is neither stretched nor compressed. The spring potential energy is then simply 0 at
y = 0, and we can define the gravitational potential energy so it is 0 at this point as well.
This means that the equilibrium position of the block will not be y = 0, and that the total
potential energy of the system at equilibrium will not automatically be Ueq = 0—unlike
in the rest of our discussion so far. We could solve for the equilibrium values and then
re-define the y coordinate and U scale to make yeq = Ueq = 0, so this is not problematic
in principle; but we do not need to do so, because the only important facts here are that
the potential energy is a minimum at the equilibrium position (as shown in Figure 1) and
that the total mechanical energy is conserved—regardless of where the zeropoints are.

If y = 0 is the initial height of the block, where the spring is “unextended,” the spring
potential energy is given for general y by
µ ¶
1 2 2 1 k 1
Uspring = mω y = m y 2 = ky 2 ,
2 2 m 2

while the gravitational potential energy is

Ugrav = mgy .

Meanwhile, the kinetic energy is always


µ ¶2
1 dy 1
K = m = mẏ 2 .
2 dt 2
–6–

(a) The total potential energy of the block at height y is


1
U (y) = Ugrav + Uspring = mgy + ky 2 .
2
The equilibrium position is that where U is minimized:
dU 1 mg
equilibrium =⇒ = 0 =⇒ mg + k × (2 y) = mg + ky = 0 =⇒ yeq = − ,
dy 2 k
as we also saw in Lecture 2. (U is a minimum, and not a maximum, at yeq because not
only is dU/dy = 0 there, but also d2 U/dy 2 = k > 0.) Note that yeq is negative because in
equilibrium the spring is stretched beyond its natural length so the block hangs below its
initial position. The absolute value |yeq | is the extension of the spring in equilibrium:

mg 0.050 kg × 9.81 m s−2


|yeq | = = ' 0.123 m .
k 4 N m−1

(b) The kinetic energy at the initial position, y = 0, is 0 (because the block is “released”
at this point), while Ugrav = 0 and Uspring = 0 there. Thus, in this set-up

Etot (y = 0) = K + Ugrav + Uspring = 0 .

When the spring is at its maximum extension, the block is at its lowest point. If we call
this y = y` (which must again be negative), then the stretch of the spring from its natural
length is |y` |. Furthermore, y` is the minimum height attained by the block, so dy/dt = 0
at this point, and the kinetic energy is again 0. Thus,
1
Etot (y = y` ) = K + Ugrav + Uspring = 0 + mgy` + ky`2 ,
2
and then conservation of energy requires
1
Etot (y` ) = Etot (y0 ) =⇒ mgy` + ky`2 = 0
2
1 2 2 mg
=⇒ ky = −mgy` =⇒ y` = − .
2 ` k
Therefore,
2 mg 2 × 0.050 kg × 9.81 m s−2
|y` | = = ' 0.245 m .
k 4 N m−1
This is exactly twice the extension of the spring when the block is in equilibrium.

(c) The kinetic energy of the block is K = 0 at y = 0, when the spring has its natural
length; and K = 0 again at y = y` = −2mg/k, when the spring has its maximum extension
for this set-up. This means that y = 0 and y = y` are the turning points of the motion;
or, put another way, the positions y = y0 and y = y` are as far from equilibrium as this
system ever gets. Moving from one of these points to the other is therefore analogous to
moving from x = +A to x = −A in our usual notation for SHM about x = 0. The time
required to do this is exactly one-half of a period:
T 1 2π π π π
t = = = = p = q = 0.351 seconds .
2 2 ω ω k/m −1
4 N m /0.050 kg
–7–

2. Simple Harmonic Motion from Conservation of Energy

In Section 1, we began by assuming simple harmonic motion and derived the potential energy
U as a function of displacement from equilibrium, then saw how the kinetic energy K behaves
and showed that the total mechanical energy, Etot = K + U , is conserved.

Now, we move some way towards a more general discussion by considering a particle of mass
m moving in one dimension under the influence of a conservative force, which happens
to be such that the potential energy of the particle at position x is quadratic in x, i.e., U ∝ x 2
in general. It is then possible to prove mathematically that, under these conditions, the
conservation of energy implies that the particle must undergo simple harmonic motion.

We are not trying to describe any specific physical system at this point; the constant of
proportionality in U ∝ x2 is just some number that we are free to represent with any symbol
or symbols that we choose. For reasons that should soon become clear, let us choose to write
the constant of proportionality as (1/2)mω 2 . Then the total mechanical energy is written
1 1
Etot = K + U = mẋ2 + mω 2 x2 . (10)
2 2
With Etot constant in time, this implies (see also Benson, Example 15.6)
dEtot 1 d ¡ 2¢ 1 d ¡ 2¢
=0 =⇒ m ẋ + mω 2 x = 0
dt 2 dt µ ¶ 2 dt µ ¶
1 dẋ 1 dx
=⇒ m 2ẋ + mω 2 2x = 0
2 dt 2 dt
(ẋ ẍ) + ω 2 (x ẋ) = 0 =⇒ ẋ ẍ + ω 2 x = 0
¡ ¢
=⇒
=⇒ ẍ + ω 2 x = 0 =⇒ ẍ = −ω 2 x ,

In the second line of this derivation, the chain rule of differentiation has been applied. In the
third line, the common factor m has been cancelled from both terms. And in the fourth line,
the common factor ẋ has been cancelled to arrive at the final result, which is the defining
equation of simple harmonic motion with angular frequency ω.

Equally, if Etot is the same at any position x, then the derivative of with respect to x is also
0, which leads to
dEtot 1 d ¡ 2¢ 1 d ¡ 2¢
=0 =⇒ m ẋ + mω 2 x = 0
dx 2 dx µ ¶ 2 dx
1 dẋ 1
=⇒ m 2ẋ + mω 2 (2x) = 0
2 dx 2
µ ¶ µ ¶
dẋ dẋ dt
=⇒ ẋ + ω 2 x = 0 =⇒ ẋ + ω2 x = 0
dx dt dx
µ ¶
1
=⇒ ẋ ẍ + ω 2 x = 0 =⇒ ẍ + ω 2 x = 0 .

(The chain rule for differentiation was applied in both the second and the third lines here.)
Thus, simple harmonic motion with an angular frequency ω follows from the conservation of
Etot as a function of position as well as from the conservation of Etot in time.

In practical terms, it is important to remember this connection between energy conservation


and SHM because in systems even slightly more complicated than a block on a spring, it can
–8–

often be easier to write down (and subsequently work with) the total mechanical energy than
it is to infer an equation of motion from a direct analysis of all the forces acting on a system.

EXAMPLE A block of mass m is attached to a vertical spring with stiffness k, via


a string that hangs over a circular pulley of mass M and radius R [moment of inertia
I = (1/2)M R2 ]. The string does not slip. Show that the angular frequency of oscillations
is given by ω 2 = 2k/(M + 2m). [see Figure 15.26 on p. 322 of Benson]

Let y = 0 be the height of the block when the spring is at its natural (zero-load) length.
(Thus, y = 0 is not the equilibrium, since the spring is extended in equilibrium; but this
definition of y = 0 is convenient in this case.)

Also define an angular coordinate θ for the wheel of the pulley, which measures the angle
(positive counterclockwise) from the vertical for a single reference point on the circumfer-
ence of the wheel. Choose this point to be the one which is precisely at the top of the
wheel (so θ = 0) when the spring is at its natural length and the block is at height y = 0.
Since the string does not slip when the block moves to any y 6= 0, it must be that y is
equal to the arclength subtending the angle θ, i.e.,
y
θ=
R
if R is the radius of the pulley wheel.

Given these definitions, for any block position y the gravitational potential energy is

Ugrav = mgy ,

while the spring potential energy is


1
Uspring = ky 2 .
2
The kinetic energy of the block is
1
Kblock = mẏ 2
2
and the kinetic energy of the rotating pulley is
1
Kpulley = I θ̇2 ,
2
for I the moment of inertia of the wheel. The total mechanical energy is therefore

Etot = Kpulley + Kblock + Ugrav + Uspring


1 2 1 1
= I θ̇ + mẏ 2 + mgy + ky 2
2 2 2
µ ¶2
1 ẏ 1 1
= I + mẏ 2 + mgy + ky 2
2 R 2 2
µ ¶
1 I 1
= m + 2 ẏ 2 + mgy + ky 2 .
2 R 2
–9–

In the third line, the definition θ = y/R has been used to write θ̇ = ẏ/R.

Now, Etot is constant in time, so


µ ¶
dEtot 1 I d 2 dy 1 d 2
= 0 =⇒ m+ 2 (ẏ ) + mg + k (y ) = 0
dt 2 R dt dt 2 dt
µ ¶
1 I 1
=⇒ m + 2 (2 ẏ ÿ) + mg ẏ + k (2 y ẏ) = 0
2 R 2
µ ¶
I k
=⇒ 1+ 2
ÿ + g + y = 0 .
mR m

In the second line, the chain rule for differentiation has been used to obtain d(ẏ 2 )/dt = 2 ẏ ÿ
and d(y 2 )/dt = 2 y ẏ. The third line follows from dividing every term in the second line by
(mẏ).

We can rearrange the last equation to solve for the acceleration ÿ, as
µ ¶ µ ¶
I k 1 k
1+ ÿ = − y − g =⇒ ÿ = − y−g
mR2 m 1 + I/(mR2 ) m
−k/m ³ mg ´
= y + .
1 + I/(mR2 ) k

As usual with this type of problem, we see that the equilibrium position, for which ÿ = 0,
is given by yeq = −mg/k. Defining the relative displacement Y to measure the distance
from equilibrium, Y ≡ (y − yeq ) = (y + mg/k), then leads to the equation of motion

k/m
Ÿ = − Y ,
1 + I/(mR2 )

which implies simple harmonic motion with an angular frequency

k/m
ω2 = .
1 + I/(mR2 )

Since the moment of inertia for the wheel is I = M R2 /2, we have I/(mR2 ) = M/(2m)
and therefore
k/m k 2k
ω2 = = = .
1 + M/(2m) m + M/2 2m + M
Lecture 4 Energy in Simple Harmonic Motion

1. Energy in Simple Harmonic Motion

We have seen that the equation for simple harmonic oscillation,

d2 x
ẍ ≡ = −ω 2 x , (1)
dt2
has the general solution
x(t) = A sin(ωt + φ0 ) (2)
for constants A (the amplitude of oscillation) and φ0 (the phase constant). This then leads to
dx
≡ ẋ = ωA cos(ωt + φ0 ) (3)
dt
d2 x
and ≡ ẍ = −ω 2 A sin(ωt + φ0 ) = −ω 2 x . (4)
dt2
We now consider the implications of these equations for the energetics of a system undergoing
mechanical oscillations, in which x is a spatial displacement from an equilibrium (at x = 0).

Consider a particle of mass m located at some position x0 relative to equilibrium in a (conser-


vative) force field, where it feels a linear restoring force as implied by equation (1):

Frestore = mẍ0 = −mω 2 x0 .

Now suppose the particle moves by an infinitesimal amount dx0 , from position x0 to position
x0 + dx0 . As it does so, the restoring force does an increment of work on the particle:

dW = Frestore dx0 = −mω 2 x0 dx0 .

Thus, the total work done by the force on the particle as it moves from equilibrium to any
position x is
x x · ¸x0 =x
1 1
Z Z
2
W = dW = − 2 0
mω x dx = − mω 2 x0 + C
0
= − mω 2 x2 .
0 0 2 x0 =0 2

(C is a constant of integration, which we do not need to know.) Now, in general, work done
on a system by a conservative force results in a change of potential energy of the system,

∆U = −W .

The minus sign here arises because if we wanted to move a particle between two positions
against the intrinsic force without acceleration, we would have to apply an external force
of exactly −Frestore to the particle, and the work −W done by that external force corresponds
–2–

to potential energy (see Chapter 8 of Benson, University Physics). In the particular case of
SHM, the formulae above mean that the difference in potential energy between equilibrium
(x = 0) and any other position in the field of a linear restoring force is
1
∆U = U (x) − Ueq = + mω 2 x2 ,
2
where the equilibrium value of the potential energy, Ueq = U (0), is just a constant number.
Re-arranging this, the potential energy at any position x is simply
1
U (x) = mω 2 x2 + Ueq . (5)
2

Because Ueq is just a constant, we are generally free to give it a value that suits our convenience
in any particular problem. In straightforward applications of SHM, it is often taken to be
Ueq = 0. In this case, a standard result is that the potential energy of a particle of mass
m in simple harmonic motion about x = 0 with angular frequency ω is
1
U = mω 2 x2 . (6)
2
Meanwhile, the kinetic energy is the usual
1
K = mẋ2 , (7)
2
since ẋ ≡ dx/dt by definition. Using equations (2) and (3) for the displacement x(t) from
equilibrium and the velocity ẋ(t), we therefore have that the potential and kinetic energies as
functions of time for a particle in SHM are
1
U = 1 2 2
2 mω x = mω 2 A2 sin2 (ω t + φ0 )
2
1
K = 1
2 mẋ
2 = mω 2 A2 cos2 (ω t + φ0 )
2
Clearly, both energies vary in time and neither is conserved on its own. However, the fact that
K varies as cos2 (ω t + φ0 ) while U varies as sin2 (ω t + φ0 ) means in effect that any change in
kinetic energy from one phase (or time) of an oscillation to another will always be balanced by
an exactly opposite change in potential energy—and vice versa (see Figure 15.9 of Benson).
That is, the total mechanical energy, Etot = K + U , is constant:
1 1
Etot = K + U = mẋ2 + mω 2 x2
2 2
1 1
= mω A cos2 (ωt + φ0 ) + mω 2 A2 sin2 (ωt + φ0 )
2 2
2 2
1
mω 2 A2 cos2 (ωt + φ0 ) + sin2 (ωt + φ0 )
£ ¤
=
2
1
=⇒ Etot = mω 2 A2 . (8)
2
In the last line of this derivation, we have used the fact that sin2 θ + cos2 θ = 1 for any angle
θ. The final value for Etot in equation (8) also requires that the amplitude A be
that of a spatial displacement, measured in units of length (it cannot be an angular
amplitude, for example; if it is not a length, then Etot will not even have the dimensions of
–3–

energy). And, again, it also assumes that we have chosen the potential-energy scale
such that U = 0 at the equilibrium x = 0.

Thus, in simple harmonic motion, the total mechanical energy depends on the square of
the angular frequency ω and on the square of the amplitude A of the oscillation. Since
both of these things are constant, energy is conserved in SHM. Conversely, conservation of
energy requires that the oscillation amplitude be constant. Note the correspondence
between equation (8) for Etot and the result that we derived in Lecture 1, for a particle on a
circular “orbit” about the origin in the (x, y) plane.

The quadratic form of the potential energy U as a function of the displacement x in equation
(6) (or, more generally, in equation [5]) is a characteristic of any simple harmonic motion. As
a result of this and the conservation of total energy, the kinetic energy can also be written as
a quadratic function of position:
1 1 1
K = Etot − U = mω 2 A2 − mω 2 x2 = mω 2 A2 − x2 ,
¡ ¢
(9)
2 2 2
which is an inverted parabola. Figure 1 illustrates the dependences of U , K, and E tot on
position in SHM (see also Figure 15.8 of Benson).

Comparing equation (6) for U and equation (8) for Etot shows that the potential energy is
equal to the total energy, U = Etot , at both x = +A and x = −A, i.e., at the endpoints of the
oscillation, when the displacement x is maximized. It must therefore be that K = E tot − U = 0
when x = ±A. (Clearly, this means that the velocity ẋ = 0 when x = ±A—which we already
knew—while we also know that the acceleration is always directed towards x = 0; thus, the
maximum and minimum displacements x = ±A are also referred to as the turning points
of the motion.) By contrast, equation (9) says that K = Etot = (1/2)mω 2 A2 at the point of
equilibrium, x = 0, where the potential energy U = (1/2)mω 2 x2 = 0.

Fig. 1.— Schematic dependence of energy on displacement from equilibrium (x = 0) for


simple harmonic motion with an amplitude of A (−A ≤ x ≤ +A). The potential energy is
U = (1/2)mω 2 x2 (solid red curve), the total energy Etot = (1/2)mω 2 A2 is constant in time and
space (solid, black, horizontal line), and the kinetic energy is K = Etot −U = (1/2)mω 2 (A2 −x2 )
(dashed blue curve). Large black dots mark the positions x where the kinetic and potential
energies are equal.
–4–

In general, then, in SHM the potential energy is a maximum when the kinetic energy
is minimum; and the potential energy is a minimum when the kinetic energy is
maximum. This “competition” between kinetic and potential energies follows from the fact
that U ∝ x2 and K ∝ ẋ2 by definition (equations [6] and [7]), plus the fact that the velocity ẋ
is 0 when the displacement x is at one of the extremes ±A while the displacement is 0 when
the velocity is at one of its extremes ±ωA. Put another way, it follows from the fact that x
and ẋ are out of phase by π/2 rad, or one-quarter of a cycle, in SHM (see the discussion of
this in Lecture 2).

Figure 1 shows explicitly that U = Etot and K = 0 at x = +A and at x = −A, while K = Etot
and U = 0 at x = 0. One complete cycle of oscillation involves the motion of a particle from,
e.g., x = +A to x = −A and then back again, so each of these endpoints is achieved once
during a full cycle; thus, U is maximized and K minimized twice per cycle. Any single
position between −A < x < +A is passed through twice per cycle (once on the way to −A,
and once on the way back from −A). This includes the equilibrium x = 0, and therefore U
is minimized and K maximized twice per cycle. The black dots in Figure 1 mark the
positions x at which the potential and kinetic energies are equal. There are two values of x
for which this occurs, and each of these two positions is passed through twice in an oscillation
cycle; thus, K = U four times per cycle. We can solve for the points in an oscillation at
which this occurs, by setting
1 1
K=U =⇒ mẋ2 = mω 2 x2 =⇒ ẋ2 = ω 2 x2 ,
2 2
which then requires (using equations [2] and [3])

ω 2 A2 cos2 (ωt + φ0 ) = ω 2 A2 sin2 (ωt + φ0 ) for K = U .

The common factor ω 2 A2 on both sides of this equation cancels, leaving


sin2 (ωt + φ0 )
cos2 (ωt + φ0 ) = sin2 (ωt + φ0 ) =⇒ tan2 (ωt + φ0 ) ≡ =1 for K = U .
cos2 (ωt + φ0 )
That is,
tan(ωt + φ0 ) = +1 or tan(ωt + φ0 ) = −1 for K = U ,
which occurs at the phases

ωt + φ0 = π/4, 3π/4, 5π/4, and 7π/4 rad


= 45◦ , 135◦ , 225◦ , and 315◦ for K = U .

Notice that the phase difference between any two consecutive points at which K = U is always
π/2 radians (= 90◦ , or one-quarter of a cycle). Since ω is the (constant) rate of change of the
phase, it takes a time π/2ω = T /4 (the period is T = 2π/ω, as usual) to move from any point
with K = U to the immediate next point with K = U . Finally, given that x = A sin(ωt + φ0 ),
the positions or displacements at which K = U are
A A
x = + √ ' +0.7071 A or x = − √ ' −0.7071 A for K = U .
2 2

The following examples show how some of the ideas above can be applied:
–5–

EXAMPLE At a certain point in the oscillation of a block+spring system the kinetic


energy is 0.1 J and the potential energy is 0.3 J. The amplitude is 20 cm and the period
is 0.8 s. Find the mass of the block and the spring constant.

The total energy is Etot = K + U = 0.1 J + 0.3 J = 0.4 J. This is conserved, and it is equal
to (1/2)mω 2 A2 . We are told that A = 0.20 m and we infer ω from the period: T = 2π/ω
so ω = 2π/T = 2π/(0.8 s) = 2.5π s−1 . Thus, the mass of the block follows from

1 2 Etot 2 × 0.4 J
Etot = mω 2 A2 =⇒ m = = = 0.324 kg .
2 ω 2 A2 (2.5π s−1 )2 (0.20 m)2
p
We also know that ω = k/m for a block on a spring, and so mω 2 = k for this system.
Therefore,
1 1 2 Etot 2 × 0.4 J
Etot = mω 2 A2 = kA2 =⇒ k = = = 20 N m−1 .
2 2 A2 (0.20 m)2

EXAMPLE A 50-gram block is attached to a vertical spring whose stiffness constant is


4 N m−1 . The block is released at the position where the spring is unextended. (a) What
is the extension of the spring when the block is in equilibrium? (b) What is the maximum
extension of the spring? (c) How long does it takes for the block to reach the lowest point?

We use considerations of energy to answer this question. Compare with the alternate,
force-based discussion in Lecture 2 of a block hanging from a vertical spring.

For this problem it is most convenient to define y = 0 as the position of the block when the
spring is neither stretched nor compressed. The spring potential energy is then simply 0 at
y = 0, and we can define the gravitational potential energy so it is 0 at this point as well.
This means that the equilibrium position of the block will not be y = 0, and that the total
potential energy of the system at equilibrium will not automatically be Ueq = 0—unlike
in the rest of our discussion so far. We could solve for the equilibrium values and then
re-define the y coordinate and U scale to make yeq = Ueq = 0, so this is not problematic
in principle; but we do not need to do so, because the only important facts here are that
the potential energy is a minimum at the equilibrium position (as shown in Figure 1) and
that the total mechanical energy is conserved—regardless of where the zeropoints are.

If y = 0 is the initial height of the block, where the spring is “unextended,” the spring
potential energy is given for general y by
µ ¶
1 2 2 1 k 1
Uspring = mω y = m y 2 = ky 2 ,
2 2 m 2

while the gravitational potential energy is

Ugrav = mgy .

Meanwhile, the kinetic energy is always


µ ¶2
1 dy 1
K = m = mẏ 2 .
2 dt 2
–6–

(a) The total potential energy of the block at height y is


1
U (y) = Ugrav + Uspring = mgy + ky 2 .
2
The equilibrium position is that where U is minimized:
dU 1 mg
equilibrium =⇒ = 0 =⇒ mg + k × (2 y) = mg + ky = 0 =⇒ yeq = − ,
dy 2 k
as we also saw in Lecture 2. (U is a minimum, and not a maximum, at yeq because not
only is dU/dy = 0 there, but also d2 U/dy 2 = k > 0.) Note that yeq is negative because in
equilibrium the spring is stretched beyond its natural length so the block hangs below its
initial position. The absolute value |yeq | is the extension of the spring in equilibrium:

mg 0.050 kg × 9.81 m s−2


|yeq | = = ' 0.123 m .
k 4 N m−1

(b) The kinetic energy at the initial position, y = 0, is 0 (because the block is “released”
at this point), while Ugrav = 0 and Uspring = 0 there. Thus, in this set-up

Etot (y = 0) = K + Ugrav + Uspring = 0 .

When the spring is at its maximum extension, the block is at its lowest point. If we call
this y = y` (which must again be negative), then the stretch of the spring from its natural
length is |y` |. Furthermore, y` is the minimum height attained by the block, so dy/dt = 0
at this point, and the kinetic energy is again 0. Thus,
1
Etot (y = y` ) = K + Ugrav + Uspring = 0 + mgy` + ky`2 ,
2
and then conservation of energy requires
1
Etot (y` ) = Etot (y0 ) =⇒ mgy` + ky`2 = 0
2
1 2 2 mg
=⇒ ky = −mgy` =⇒ y` = − .
2 ` k
Therefore,
2 mg 2 × 0.050 kg × 9.81 m s−2
|y` | = = ' 0.245 m .
k 4 N m−1
This is exactly twice the extension of the spring when the block is in equilibrium.

(c) The kinetic energy of the block is K = 0 at y = 0, when the spring has its natural
length; and K = 0 again at y = y` = −2mg/k, when the spring has its maximum extension
for this set-up. This means that y = 0 and y = y` are the turning points of the motion;
or, put another way, the positions y = y0 and y = y` are as far from equilibrium as this
system ever gets. Moving from one of these points to the other is therefore analogous to
moving from x = +A to x = −A in our usual notation for SHM about x = 0. The time
required to do this is exactly one-half of a period:
T 1 2π π π π
t = = = = p = q = 0.351 seconds .
2 2 ω ω k/m −1
4 N m /0.050 kg
–7–

2. Simple Harmonic Motion from Conservation of Energy

In Section 1, we began by assuming simple harmonic motion and derived the potential energy
U as a function of displacement from equilibrium, then saw how the kinetic energy K behaves
and showed that the total mechanical energy, Etot = K + U , is conserved.

Now, we move some way towards a more general discussion by considering a particle of mass
m moving in one dimension under the influence of a conservative force, which happens
to be such that the potential energy of the particle at position x is quadratic in x, i.e., U ∝ x 2
in general. It is then possible to prove mathematically that, under these conditions, the
conservation of energy implies that the particle must undergo simple harmonic motion.

We are not trying to describe any specific physical system at this point; the constant of
proportionality in U ∝ x2 is just some number that we are free to represent with any symbol
or symbols that we choose. For reasons that should soon become clear, let us choose to write
the constant of proportionality as (1/2)mω 2 . Then the total mechanical energy is written
1 1
Etot = K + U = mẋ2 + mω 2 x2 . (10)
2 2
With Etot constant in time, this implies (see also Benson, Example 15.6)
dEtot 1 d ¡ 2¢ 1 d ¡ 2¢
=0 =⇒ m ẋ + mω 2 x = 0
dt 2 dt µ ¶ 2 dt µ ¶
1 dẋ 1 dx
=⇒ m 2ẋ + mω 2 2x = 0
2 dt 2 dt
(ẋ ẍ) + ω 2 (x ẋ) = 0 =⇒ ẋ ẍ + ω 2 x = 0
¡ ¢
=⇒
=⇒ ẍ + ω 2 x = 0 =⇒ ẍ = −ω 2 x ,

In the second line of this derivation, the chain rule of differentiation has been applied. In the
third line, the common factor m has been cancelled from both terms. And in the fourth line,
the common factor ẋ has been cancelled to arrive at the final result, which is the defining
equation of simple harmonic motion with angular frequency ω.

Equally, if Etot is the same at any position x, then the derivative of with respect to x is also
0, which leads to
dEtot 1 d ¡ 2¢ 1 d ¡ 2¢
=0 =⇒ m ẋ + mω 2 x = 0
dx 2 dx µ ¶ 2 dx
1 dẋ 1
=⇒ m 2ẋ + mω 2 (2x) = 0
2 dx 2
µ ¶ µ ¶
dẋ dẋ dt
=⇒ ẋ + ω 2 x = 0 =⇒ ẋ + ω2 x = 0
dx dt dx
µ ¶
1
=⇒ ẋ ẍ + ω 2 x = 0 =⇒ ẍ + ω 2 x = 0 .

(The chain rule for differentiation was applied in both the second and the third lines here.)
Thus, simple harmonic motion with an angular frequency ω follows from the conservation of
Etot as a function of position as well as from the conservation of Etot in time.

In practical terms, it is important to remember this connection between energy conservation


and SHM because in systems even slightly more complicated than a block on a spring, it can
–8–

often be easier to write down (and subsequently work with) the total mechanical energy than
it is to infer an equation of motion from a direct analysis of all the forces acting on a system.

EXAMPLE A block of mass m is attached to a vertical spring with stiffness k, via


a string that hangs over a circular pulley of mass M and radius R [moment of inertia
I = (1/2)M R2 ]. The string does not slip. Show that the angular frequency of oscillations
is given by ω 2 = 2k/(M + 2m). [see Figure 15.26 on p. 322 of Benson]

Let y = 0 be the height of the block when the spring is at its natural (zero-load) length.
(Thus, y = 0 is not the equilibrium, since the spring is extended in equilibrium; but this
definition of y = 0 is convenient in this case.)

Also define an angular coordinate θ for the wheel of the pulley, which measures the angle
(positive counterclockwise) from the vertical for a single reference point on the circumfer-
ence of the wheel. Choose this point to be the one which is precisely at the top of the
wheel (so θ = 0) when the spring is at its natural length and the block is at height y = 0.
Since the string does not slip when the block moves to any y 6= 0, it must be that y is
equal to the arclength subtending the angle θ, i.e.,
y
θ=
R
if R is the radius of the pulley wheel.

Given these definitions, for any block position y the gravitational potential energy is

Ugrav = mgy ,

while the spring potential energy is


1
Uspring = ky 2 .
2
The kinetic energy of the block is
1
Kblock = mẏ 2
2
and the kinetic energy of the rotating pulley is
1
Kpulley = I θ̇2 ,
2
for I the moment of inertia of the wheel. The total mechanical energy is therefore

Etot = Kpulley + Kblock + Ugrav + Uspring


1 2 1 1
= I θ̇ + mẏ 2 + mgy + ky 2
2 2 2
µ ¶2
1 ẏ 1 1
= I + mẏ 2 + mgy + ky 2
2 R 2 2
µ ¶
1 I 1
= m + 2 ẏ 2 + mgy + ky 2 .
2 R 2
–9–

In the third line, the definition θ = y/R has been used to write θ̇ = ẏ/R.

Now, Etot is constant in time, so


µ ¶
dEtot 1 I d 2 dy 1 d 2
= 0 =⇒ m+ 2 (ẏ ) + mg + k (y ) = 0
dt 2 R dt dt 2 dt
µ ¶
1 I 1
=⇒ m + 2 (2 ẏ ÿ) + mg ẏ + k (2 y ẏ) = 0
2 R 2
µ ¶
I k
=⇒ 1+ 2
ÿ + g + y = 0 .
mR m

In the second line, the chain rule for differentiation has been used to obtain d(ẏ 2 )/dt = 2 ẏ ÿ
and d(y 2 )/dt = 2 y ẏ. The third line follows from dividing every term in the second line by
(mẏ).

We can rearrange the last equation to solve for the acceleration ÿ, as
µ ¶ µ ¶
I k 1 k
1+ ÿ = − y − g =⇒ ÿ = − y−g
mR2 m 1 + I/(mR2 ) m
−k/m ³ mg ´
= y + .
1 + I/(mR2 ) k

As usual with this type of problem, we see that the equilibrium position, for which ÿ = 0,
is given by yeq = −mg/k. Defining the relative displacement Y to measure the distance
from equilibrium, Y ≡ (y − yeq ) = (y + mg/k), then leads to the equation of motion

k/m
Ÿ = − Y ,
1 + I/(mR2 )

which implies simple harmonic motion with an angular frequency

k/m
ω2 = .
1 + I/(mR2 )

Since the moment of inertia for the wheel is I = M R2 /2, we have I/(mR2 ) = M/(2m)
and therefore
k/m k 2k
ω2 = = = .
1 + M/(2m) m + M/2 2m + M
Lectures 6/7 Damped Oscillation

To move beyond simple harmonic motion and consider the behaviour of more realistic (and
hence complicated) systems—which will in general be subject to frictional energy losses, for
example, and may additionally be acted upon by external forces—it is often convenient to
work with complex numbers and functions. Moreover, the exponential function (both
real and complex) appears frequently in the analysis of oscillating systems. Thus, the first
two sections of these notes review a few main results pertaining to complex numbers and the
exponential function. Section 1, in particular, should be read alongside the original Maths
Notes on Complex Numbers from Semester 1 (PHY-10010). After this brief mathematics
revision, we discuss damped oscillators in Sections 3 and 4 of these notes.

1. Complex Numbers: Important Results

All of the mathematical treatment of oscillations and waves could be done in terms of complex
numbers and complex functions. Although we have not yet worked in terms of complex numbers
at all—and we will not overly stress this approach in this module—it can be more convenient in
some applications to do so. Moreover, in some more advanced topics (e.g., quantum mechanics),
complex analysis is indispensable to the physics.

There are four main facts about complex numbers to be kept in mind:

(1) The definition of i is



i ≡ −1
which of course is the same as
i2 = −1
and also leads to (dividing both sides of the above by i)
1 1
i = − or = −i
i i

(2) The standard form of a complex number is

z = a + ib ,

in which a and b are both real numbers. These are the real part and the imaginary
part of z: )
Re(z) = a
both real
Im(z) = b

(3) The complex conjugate of a complex number, in standard form, is

z∗ ≡ a − i b
–2–

In general, the conjugate is obtained simply by replacing i with −i wherever it is found in


the original complex number. The sum of a complex number and its conjugate is
always real:
z + z ∗ = (a + i b) + (a − i b) = 2 a = 2 Re(z)

(4) The complex exponential function is defined by

ei θ = cos θ + i sin θ

where θ is any number. The conjugate of the complex exponential is thus

e−i θ = cos θ − i sin θ

Most importantly for us, the complex exponential of a function is defined in precisely
the same way: for example, with f (t) some arbitrary function of time,

ei f (t) = cos [f (t)] + i sin [f (t)]

and similarly for the conjugate complex function, e−i f (t) .

1.1. Complex exponentials in SHM

The presence of the sine and cosine functions in the complex exponential suggest a connection
with simple harmonic motion, in which the displacement from equilibrium is a sinusoidal
function of time. It is a matter of straightforward algebra to make this connection explicit.

We know that the general solution to the defining equation of SHM, ẍ = −ω 2 x, is

x(t) = A sin(ω t + φ0 )

where A and φ0 are constants related to initial conditions, and ω is an intrinsic property of
the system under study. We can re-write this solution by expanding the sine of the sum of the
angles ω t and φ0 (cf. the Maths Handbook):

sin (ω t + φ0 ) = sin(ω t) cos φ0 + cos(ω t) sin φ0

so that

x(t) = A sin(ω t + φ0 )
= A sin(ω t) cos(φ0 ) + A cos(ω t) sin(φ0 )
= [A sin φ0 ] cos(ω t) + [A cos φ0 ] sin(ω t)

But now, we also have that

eiωt = cos(ω t) + i sin(ω t) (1)


−iωt
e = cos(ω t) − i sin(ω t) (2)

so we can re-write cos(ω t) and sin(ω t) in terms of complex exponentials.


–3–

Adding equations (1) and (2) gives


1 ¡ iωt
eiωt + e−iωt = 2 cos(ω t) e + e−iωt
¢
=⇒ cos(ω t) =
2
while subtracting equation (2) from equation (1) yields

1 ¡ iωt i¡
eiωt − e−iωt = 2i sin(ω t) e − e−iωt = − eiωt − e−iωt
¢ ¢
=⇒ sin(ω t) =
2i 2
The last equality has made use of the fact that 1/i = −i in general, so 1/(2i) = −i/2.

We can now use these expressions for cos(ω t) and sin(ω t) to write the displacement x(t) in
SHM as

x(t) = [A sin φ0 ] cos(ω t) + [A cos φ0 ] sin(ω t)


· ¸ · ¸
1 ¡ iωt −i ¡ iωt
e + e−iωt e − e−iωt
¢ ¢
= [A sin φ0 ] + [A cos φ0 ]
2 2
µ ¶ µ ¶
1 i iωt 1 i
= A sin φ0 − A cos φ0 e + A sin φ0 + A cos φ0 e−iωt
2 2 2 2
≡ B eiωt + B ∗ e−iωt

in which B ≡ (A/2) sin φ0 − i (A/2) cos φ0 is a complex constant which incorporates the two
original constants A and φ0 in its real and imaginary parts. Notice that the B ∗ multiplying
e−iωt here is exactly the complex conjugate of the B multiplying eiωt . Thus, x(t) as written
here is the sum of a complex function (B eiωt ) and its complex conjugate, and so ultimately
x(t) is still a real function of time.

We have just shown that the functions

x(t) = A sin(ω t + φ0 )

and
x(t) = B eiωt + B ∗ e−iωt
are equivalent ways of writing the general solution to the equation of motion, ẍ = −ω 2 x,
that defines simple harmonic oscillation. The two forms of the solution are connected through
their constants: B = (A/2) sin φ0 − i (A/2) cos φ0 .

2. Derivatives of the Exponential Function

As we will see in the next Section, the exponential function (whether real or complex) comes up
naturally in the solution to the equation of motion for a damped oscillator. Here we recall the
fact that all derivatives of the function eαt, where α is a numerical constant (either
real or complex), are proportional to the function itself. That is, if we define either
u(t) ≡ B e+αt or u(t) ≡ B e−αt , with both B and α constants (and both possibly complex),
the first several derivatives of u with respect to t are (use the chain rule of differentiation to
–4–

recover these for yourself):

u = B e+αt B e−αt
du/dt = +α B e+αt = α u(t) −α B e−αt = −α u(t)
d2 u/dt2 = +α2 B e+αt = α2 u(t) +α2 B e−αt = +α2 u(t)
d3 u/dt3 = +α3 B e+αt = α3 u(t) −α3 B e−αt = −α3 u(t)
etc . . .

Notice the second time derivative here in particular: if the constant α is imaginary—so, say,

either α = i ω or α = −i ω, where i = −1 as usual and ω > 0 is some real constant—then
α2 = −ω 2 , and so d2 u/dt2 = −ω 2 u. This is the equation for simple harmonic motion at
angular frequency ω, and therefore this shows again that the complex exponential function
eiωt and the conjugate function e−iωt form a valid basis for expressing the general solution for
the time-dependent displacement of a simple harmonic oscillator.

3. Damped Oscillations: General Equation and Solution

With the mathematical revision above, we are prepared now to consider equations of motion
for (would-be) oscillators that are subject to forces in addition to a restoring force that by
itself would lead to simple harmonic motion.

We have been concentrating to this point on the case that the net force on an object or a
system is a restoring force with a magnitude that is either exactly or approximately (as in near
a stable equilibrium) proportional to the displacement from equilibrium. Thus, we have been
considering the equation of motion
ẍ = −ω 2 x , (3)
or, in terms of forces,
Frestore = mẍ = −mω 2 x .
We have seen that the solution to this equation of motion is

x(t) = A sin(ωt + φ0 ) ,

or equivalently (from Section 1.1 above),

x(t) = B eiωt + B ∗ e−iωt

with the complex constant B = (A/2) sin φ0 − i (A/2) cos φ0 . We have also seen that the total
mechanical energy (sum of kinetic and potential energies) is conserved in a system governed by
equation (3); and that Etot = 12 mω 2 A2 , so the conservation of mechanical energy is equivalent
to a constant amplitude A of oscillation (when x and therefore A are spatial displacements with
units of length).

In reality, of course, the total mechanical energy of any system in motion is never perfectly
conserved. There is always some loss of mechanical energy—perhaps due to the effects of, e.g.,
internal friction, air resistance, fluid viscosity, and so on—which is then dissipated (as heat, for
example). Thus, while an understanding of simple harmonic oscillations, with conservation of
–5–

mechanical energy and constant amplitude, is an important starting point for many physical
analyses, the simplest version of the problem that we have started with is necessarily an
idealization.

In the context of simple harmonic motion, we refer generically to a resistive force, which works
to dissipate mechanical energy from a system in motion, as a damping force. Such forces are
generally velocity-dependent (e.g., faster motion through air implies greater air resistance).
The simplest possible dependence is one in which the magnitude of the damping force is just
directly proportional to the speed |ẋ|, and in fact it does happen that real damping forces can
often be approximated reasonably well in this way (see Benson, University Physics, Chapter
6.4). Thus, let us now consider
Fdamp = −γ ẋ .
This defines the damping constant γ. Note the SI units of γ: F has units of N and ẋ has
units of m s−1 , so the units of γ are N/(m s−1 ) = (kg m s−2 )/(m s−1 ) = kg s−1 .

The minus sign in Fdamp = −γ ẋ indicates that the damping force always opposes the instan-
taneous motion. This is slightly different from Frestore = −mω 2 x for the force in pure simple
harmonic motion: in that case the restoring force is always directed oppositely to the instanta-
neous displacement from equilibrium, but the force (acceleration) and the velocity will be in the
same direction whenever the object is already moving towards the equilibrium position—that
is, for half of every cycle.

The net force on a would-be oscillator subjected to a linear damping force that is linear in
velocity is simply the sum

Fnet = Fdamp + Frestore =⇒ mẍ = −γ ẋ − mω02 x .

The subscript “0” on ω0 has been introduced here to denote the angular frequency that would
obtain in the absence of any damping of the system; it does not refer to an initial value for
ω0 , since this is still just a constant. Thus, ω0 is referred to as the natural angular
frequency (or the undamped angular frequency) of a system.

Bringing all terms to the left-hand side in the above equation and dividing through by the
mass m leads to the basic differential equation governing the motion of a damped oscillator:
γ
ẍ + ẋ + ω02 x = 0 . (4)
m
Of course, if γ = 0 (no damping) this reduces exactly to the defining equation of simple
harmonic oscillation (ẍ + ω02 x = 0, or ẍ = −ω02 x), which we have been working with thus far.

How does the displacement x from equilibrium behave as a function of time when damping
is present? Obviously, the behaviour that satisfies equation (4) will be more general—which
is to say, more complicated—than the simple sinusoidal dependence in x(t) that obtains for
the familiar, undamped equation of motion (3). In the rest of this section we work step-
by-step through the procedure to solve our new equation (4), but you should cross-reference
this work with the concurrent Maths lectures on Differential Equations. In particular, note
that equation (4) is an example of a second-order, linear, homoegeneous, ordinary
differential equation. All of these adjectives are defined in the Maths lectures, and our
solution of the equation is an application of the techniques in the Maths lectures.
–6–

Because all of the coefficients of ẍ, ẋ, and the function x itself in equation (4) are strictly
constant numbers (involving γ, m, and ω0 , which are all intrinsic properties of the particular
system under investigation), it proves useful now to recall the properties of the exponential
function (real or complex) as revised in Section 2 above. That is, we know that every time-
derivative of the function e±αt , for α a constant (which may be complex), is just proportional to
the original function e±αt . Given this, and seeing that every term in the fundamental equation
of motion (4) involves some constant multiplying one of x, ẋ, or ẍ, we can guess that it might
be helpful to look for a solution x(t) that is itself proportional to an exponential of the form
eαt (with α needing to be determined self-consistently so that the differential equation [4] is
in fact satisfied). If this is the case, then every term on the left-hand side of equation (4) will
ultimately contain a factor eαt , which will then cancel out (because the right-hand side of the
equation is 0) and leave a new equation that is equivalent to the original but hopefully easier
to solve directly.

Let us see in detail how this works. We stipulate that the solution x(t) to equation (4) for a
damped oscillator is of the form eαt times another function of t, which we call X(t):

x(t) ≡ eαt X(t) . (5)

This is completely general: we can always re-write any arbitrary function in this way, as
the product of two other functions. The benefit of doing so here will become clear shortly.
However, notice one implication for our new function X(t): it cannot be an exponential, since
by writing x(t) = eαt X(t) we mean to isolate (or factor out) any and all exponential time-
dependence from the full solution x(t), and then to deduce what (if any) other form of time
dependence might be left in the new function X(t).

Starting from the definition in equation (5), we can use the chain rule and the product rule of
differentiation together to find the derivatives ẋ and ẍ in terms of the derivatives of the new
function X(t), without having to specify the constant α quite yet:
dx d £ αt ¤
ẋ ≡ = e X
dt ·dt ¸ · ¸
d αt αt dX
= e X +e
dt dt
= α eαt X + eαt Ẋ , (6)

and then
d2 x d
ẍ ≡ = ẋ
dt2 dt
d h pt i
= α e X + eαt Ẋ
dt
d £ αt ¤ d h αt i
= αe X + e Ẋ
dt dt
·µ ¶ µ ¶¸ "µ ¶ à !#
d αt αt dX d αt αt dẊ
= αe X +αe + e Ẋ + e
dt dt dt dt
h i h i
= α2 eαt X + α eαt Ẋ + α eαt Ẋ + eαt Ẍ
= α2 eαt X + 2α eαt Ẋ + eαt Ẍ . (7)
–7–

Now we can use equations (5), (6), and (7) to substitute for x, ẋ, and ẍ in equation (4), hence
writing it in terms of eαt and the new function X(t). That is, the original equation
γ
ẍ + ẋ + ω02 x = 0
m
becomes, with the stipulation x(t) ≡ eαt X(t),

eαt Ẍ + 2α eαt Ẋ + α2 eαt X [ẍ]


γ γ γ
+ eαt Ẋ + α eαt X [+ ẋ]
m m m
+ω02 eαt X = 0 [+ω02 x]

which simplifies to
h ³ γ´ ³ γ ´ i
eαt Ẍ + 2α + Ẋ + α2 + α + ω02 X = 0 . (8)
m m

As expected, the factor eαt can now be cancelled from the equation, leaving an equation for the
new function X(t) only—inside the square brackets of equation (8)—which must itself always
equal 0 (because eαt never does!). Moreover, since we introduced the definition x(t) = eαt X(t)
in the first place entirely for our own convenience, we are free now to specify the constant α
(in terms of the original, physical parameters of the problem, which again are γ, m, and ω 0 )
in any way that makes it easier for us to go further in obtaining the full solution x(t).

Evidently, if we choose α such that


γ
α2 + m
α + ω02 = 0 (9)

then the term in X(t) vanishes from inside the brackets in equation (8), leaving us with
an equation that involves only Ẍ and Ẋ and is easier to solve in general:
³ γ´
Ẍ + 2α + Ẋ = 0 .
m
The proviso is that the quadratic equation (9) for α means that we must have
" r #
1 γ γ2
α = − ± − 4ω02
2 m m2
q
=⇒ α = −γ/(2m) ± γ 2 /(4 m2 ) − ω02 (10)

This gives us up to two possible values for the constant α in the exponential part of x(t) =
eαt X(t), and further implies
r
γ γ2
2α + = ±2 − ω02 .
m 4 m2
Therefore, the equation left to solve for X(t) is in fact just
³ γ´
Ẍ + 2α + Ẋ = 0
· q m¸
=⇒ Ẍ ± 2 γ 2 /(4 m2 ) − ω02 Ẋ = 0 (11)
–8–

Again, by construction the solution to this equation for X(t) is multiplied by eαt to obtain the
full solution to the damped equation of motion:

x(t) = eαt X(t) (12)

There are three different types of motion x(t), depending on whether the properties of the
oscillator in question happen to make the quantity γ 2 /(4m2 ) − ω02 —appearing under the
£ ¤

square root in equations (10) and (11)—be zero, positive, or negative.

£ ¤
• CASE I γ 2 /(4m2 ) − ω02 = 0 =⇒ γ = 2mω0 Critical Damping
When the damping constant and the natural angular frequency of the undamped oscil-
lator have the very specific relationship γ = 2mω0 , the motion of the oscillator is said to
be critically damped.
Equation (10) for the value of the constant α reduces in this case to a single number:
γ
α = − ,
2m
which is also
α = −ω0
since γ = 2mω0 . Moreover, the equation for our subsidiary function, X(t), becomes
particularly simple:
Ẍ = 0 ,
p
since the constant multiplying Ẋ in equation (11) is 2 γ 2 /(4m2 ) − ω02 = 0 in this case.
We can integrate the equation Ẍ = 0 immediately (twice):

d
Ẍ = Ẋ = 0 =⇒ Ẋ = C1 ,
dt
where C1 is a constant of integration, and then
Z Z
Ẋ = C1 =⇒ dX/dt = C1 =⇒ dX = C1 dt
=⇒ X(t) = C1 t + C2 ,

where C2 is another constant of integration, which is independent of C1 . There-


fore, the full solution for the displacement of a critically damped oscillator is (from
eq. [12])

[γ = 2mω0 ] : x(t) = eαt X(t) = e−γ t/(2m) [C1 t + C2 ]


= e−ω0 t [C1 t + C2 ] . (13)

The behaviour of this x(t) is illustrated in Section 4 below. As discussed further there,
it is a non-oscillatory, aperiodic motion, in which the displacement approaches 0
(equilibrium) exponentially at late times. Critical damping allows for the fastest
possible return of a displaced oscillator to equilibrium.
–9–

£ ¤
• CASE II γ 2 /(4m2 ) − ω02 > 0 =⇒ γ > 2mω0 Overdamping
When the damping constant is large enough that γ > 2mω0 , the oscillator is said to be
overdamped or (interchangeably) heavily damped.
In this case, we define a new constant,

γ2
q2 ≡ − ω02 > 0
4m2
so that equation (10) gives two possible values for the constant α:
r
γ γ2 γ
α1 = − + 2
− ω02 = − +q
2m r 4m 2m
γ γ2 γ
α2 = − − 2
− ω02 = − −q
2m 4m 2m
By convention, q itself is positive; and it is real, since γ > 2mω0 and therefore q 2 > 0.
Meanwhile, equation (11) says that the sum of the second time-derivative of X(t) plus a
non-zero constant times the first time-derivative must always be exactly 0. The only way
to have this—since X(t) is by construction not an exponential function—is to have both
derivatives Ẋ = dX/dt and Ẍ = d2 X/dt2 separately being 0 at all times. That is, we
require in this case that X(t) is itself just a constant: say X(t) ≡ B, a real number.
The full displacement x(t) then may equally well be either one of

x1 (t) = eα1 t X(t) = B eα1 t = B e−γt/(2m)+qt


x2 (t) = eα2 t X(t) = B eα2 t = B e−γt/(2m)−qt

Thus, the most general solution for x(t) is actually an arbitrary linear combination of
x1 (t) and x2 (t):

x(t) = B1 eα1 t + B2 eα2 t = B1 e−γt/(2m) e+qt + B2 e−γt/(2m) e−qt

in which B1 and B2 are two independent real constants, which will be different in
general.
In summary, then, for overdamping we have
r
−γ t/(2m) qt −qt γ2
− ω02 > 0 . (14)
£ ¤
[γ > 2mω0 ] : x(t) = e B1 e + B 2 e ; q≡
4m2

The behaviour of this x(t) is illustrated in Section 4 below. For the moment, note that
there is no trace of sinusoidal variations in x; overdamping leads to non-oscillatory,
aperiodic motion in which a displaced oscillator approaches equilibrium (x = 0) expo-
nentially at late times. This return to equilibrium is always slower for an overdamped os-
cillator (γ > 2mω0 ) than for an otherwise identical system that is only critically damped
(γ = 2mω0 ).
– 10 –

£ ¤
• CASE III γ 2 /(4m2 ) − ω02 < 0 =⇒ γ < 2mω0 Underdamping
When the damping constant is so small that γ < 2mω0 , the oscillator is said to be
underdamped or (interchangeably) lightly damped.
In this case, we define a new positive constant,

γ2
ω 2 ≡ ω02 − > 0 (15)
4m2
so that ω is positive and real, and the square root appearing in equations (10) and (11)
is the imaginary number,
r
γ2 p
2
− ω02 = (−1) × ω 2 = i ω ,
4m
Equation (10) in particular thus gives two possible complex values for the constant α:
r
γ γ2 γ
α1 = − + 2
− ω02 = − + iω
2m r 4m 2m
γ γ2 γ
α2 = − − 2
− ω02 = − − iω
2m 4m 2m
Notice that α2 is the complex conjugate of α1 .
As in the case of overdamping, equation (11) in the present, underdamped case again
implies that each of the time derivatives Ẋ and Ẍ must be identically 0, and thus that
X(t) = constant. If we call this constant B again, but allow now for the fact that B is
complex because α is complex, then the displacement of an underdamped oscillator may
be either one of

x1 (t) = eα1 t X(t) = B eα1 t = B e−γt/(2m)+i ωt


x2 (t) = eα2 t X(t) = B eα2 t = B e−γt/(2m)−i ωt

Thus, the most general solution for x(t) is again a linear combination of x 1 (t) and x2 (t):

x(t) = B1 eα1 t + B2 eα2 t

However, now the complex constants B1 and B2 cannot be independent: because we


require x(t) to be a real function of time, and we have already that α2 is the complex
conjugate of α1 , B2 must be the complex conjugate of B1 . Ultimately, then,

x(t) = B1 e−γt/(2m) e+iωt + B1∗ e−γt/(2m) e−iωt


= e−γt/(2m) B1 eiωt + B1∗ e−iωt
£ ¤

= e−γt/(2m) [A0 sin(ω t + φ0 )]

The last line, in which the complex number B1 has given way to two independent real
constants A0 and φ0 , follows from our discussion in Section 1.1 above.
To summarize, for an underdamped oscillator we have
r
γ2
[γ < 2mω0 ] : x(t) = A0 e−γt/(2m) sin(ωt + φ0 ) ; ω≡ ω02 − > 0. (16)
4m2
– 11 –

The behaviour of this x(t) is also illustrated in Section 4. It can be seen already, however,
that the displacement oscillates about the equilibrium (x = 0) with a period T = 2π/ω
for ω as given in equation (16). The damping force in the underdamped case works
essentially to make the amplitude of the oscillation decrease steadily over time: friction or
any similar force removes mechanical energy from a system, and we know that E tot ∝ A2 ,
so A must decrease along with Etot .

Before going on to discuss the qualitative features of x(t) for underdamped, overdamped, and
critically damped oscillators , it is worth noting a couple of general mathematical points about
the formulae we have just derived.

First, recall that to obtain the solutions we first defined a constant α and a function X(t) such
that x(t) ≡ eαt X(t); and then, after some algebra, we stipulated α2 + (γ/m)α + ω02 = 0
(equation [9]) to facilitate a complete solution. This equation appears in the Maths lectures on
Differential Equations, where it is referred to as the auxiliary equation. Our development
here is an independent justification of this equation, which is of a type that appears frequently
in the analysis of various second-order differential equations.

Second, notice that the solutions for x(t) in equation (13) for critically damped oscillators,
in equation (14) for overdamping, and in equation (16) for underdamping all involve two
independent, real constants: C1 and C2 for critical damping; B1 and B2 for overdamping;
and A and φ0 for underdamping. These always come from outside the original equation
of motion (where the only constants appearing are fundamental properties of the physical
system: m, γ, and ω0 ); essentially, they are all constants of integration. There are exactly two
independent constants in each case because the underlying differential equation being solved
is a second-order equation. The values of the constants need to be inferred in any specific
application from two independent pieces of specific information about the state of a particular
system at some particular time. Often (though not always) this additional information will
come from initial conditions—for example, the values of x and ẋ at time t = 0.

4. Damped Oscillations: Physical Discussion

In this section we discuss the qualitative features of the solutions derived above for the dis-
placement of underdamped, overdamped, and critically damped oscillators. We do this in the
order opposite to which we obtained the relevant equations in Section 3.

Underdamping (γ < 2mω0 )

Figure 1 shows, first (in the top half), the displacement versus time of an undamped simple
harmonic oscillator (γ = 0) with an assumed phase constant φ0 = 0, amplitude A0 and natural
angular frequency ω0 . That is, x(t) = A0 sin(ω0 t), the usual solution to the usual equation of
motion ẍ = −ω02 x for SHM. Vertical ticks along the time axis mark the times t = T0 , t = 2T0 ,
and t = 3T0 , where T0 is the period of (undamped) oscillation: T0 = 2π/ω0 , as always.
– 12 –

Fig. 1.— Top: motion of a simple harmonic oscillator with no damping. The phase constant
in this example is φ0 = 0, so x(t) = A0 sin(ω0 t). Bottom: underdamped (lightly damped)
oscillator with the same natural amplitude A0 and natural angular frequency ω0 as in the top
(and phase constant φ0 = 0 again). The damping constant is γ = 18 mω0 for this example.

The lower half of Figure 1 shows the motion of the same oscillator (i.e., same φ 0 = 0, so x = 0
at t = 0; and the same natural amplitude A0 and natural angular frequency ω0 ) when it is
lightly damped or underdamped by a force with damping constant γ < 2mω0 . Repeating
the general solution from Section 3 for this case (equation [16]),
r
γ2
γ < 2mω0 : x(t) = A0 e−γt/(2m) sin(ωt + φ0 ) ω = ω02 − > 0
4m2
The solid curve in the lower half of Figure 1 is this equation with γ = (1/8)mω 0 chosen for
illustrative purposes.

This formula for x(t) consists of an oscillatory part, sin(ω t + φ0 ) as usual, multiplied by a
function that acts as a time-dependent amplitude,
µ ¶ µ ¶
γt t
A(t) ≡ A0 exp − = A0 exp − . (17)
2m 2m/γ
The presence of the oscillatory term in x(t) meansp
that the motion of an underdamped oscillator
is still periodic: given the angular frequency ω = ω02 − γ 2 /(4m2 ) as just above, the associated
– 13 –

period is
2π 2π
T = =q . (18)
ω ω02 − γ 2 /(4m2 )
Evidently, because the damped angular frequency is always lower than the natural
angular frequency of the undamped system (the equation for ω gives ω < ω0 for any γ >
0), the oscillation period of an underdamped oscillator is always longer than the
natural period of the undamped system. For the specific example shown in Figure
p
1, the difference is relatively small: T = 256/255 T0 , as can be verified by substituting
γ = (1/8)mω0 into equation (18). The ticks along the time axis in the bottom part of Figure
1 mark the times t = T , t = 2T , and t = 3T .

What is more evident in Figure 1 is the steady decay of the oscillation amplitude. This
occurs generically in any underdamped system: because of the negative-exponential function
of time in A(t) ≡ A0 e−γt/(2m) , which multiplies the sinusoidal part of x(t), the amplitude can
only ever decrease (from a maximum value of A0 at time t = 0) as time increases.

At any given instant, the value of sin(ω t + φ0 ) must lie between ±1, and so, because x(t) =
A(t) sin(ω t+φ0 ), the displacement must always lie between the curves xmax (t) = +A0 e−γt/(2m)
and xmin (t) = −A0 e−γt/(2m) . Put another way, the excursions of an underdamped oscillator
about the equilibrium x = 0 are contained within an envelope defined by the curves x = ±A(t).
These curves are drawn in Figure 1 for our current example of γ = (1/8)mω 0 . The oscillator
hits the upper envelope exactly once every cycle and it hits the lower envelope exactly once
every cycle—in essentially the same way as x(t) for an undamped oscillator hits a constant
maximum of +A0 once every cycle and a constant minimum of −A0 once every cycle.

The decay of amplitude in underdamped oscillation means that x(t) approaches the equilib-
rium position x = 0 in the long term, but it does so somewhat slowly, by repeatedly
overshooting the equilibrium and subsequently over-correcting.

Because the total mechanical energy Etot is proportional to the square of oscillation amplitude
in simple harmonic motion, we expect—and it can be shown rigorously—that for underdamped
oscillations the mechanical energy decays in time as the square of the decaying amplitude:
· µ ¶¸2 µ ¶ µ ¶
2 2 γt γt t
Etot (t) ∝ A(t) = A0 exp − ∝ exp − = exp − . (19)
2m m m/γ
This steady loss (dissipation) of energy is caused directly by the damping force,
which generally converts mechanical energy into heat (think of friction or air resistance).

Equations (17) and (19) also indicate the characteristic timescales for decay of the
amplitude and the mechanical energy. That is, if the instantaneous value of A(t) is
measured at any particular time, it will fall to be 1/e times this value after a time

τA = 2m/γ .

Similarly, the time it takes for the mechanical energy to fall by a factor of e is

τE = m/γ ,

which is exactly one-half of the timescale for amplitude decay.


– 14 –

Fig. 2.— Motion of an overdamped oscillator with natural angular frequency ω 0 . For a given
system (fixed natural amplitude A0 and natural angular frequency ω0 ) and identical initial
conditions (fixed x = 0 at t = 0 and fixed non-zero velocity ẋ at t = 0 in this example),
increasing the damping constant (γ = 4mω0 to γ = 8mω0 in this plot) leads to a smaller
maximum displacement from the equilibrium position x = 0, and a slower return towards
equilibrium. Ticks on the time axis are at t = T0 and t = 2 T0 , the first two periods of natural
oscillation for the undamped system.

Overdamping (γ > 2mω0 )

To repeat the solution obtained in Section 3 for the displacement of overdamped or heavily
damped oscillators (equation [14]):
r
−γ t/(2m) qt −qt γ2
− ω02 > 0
£ ¤
γ > 2mω0 : x(t) = e B1 e + B 2 e q≡
4m2

Figure 2 shows such x(t) for two different values of the damping constant: γ = 4mω 0 and
a higher γ = 8mω0 . Both curves are obtained by choosing constants B1 and B2 such that
the oscillator has the same position at t = 0 (specifically, x = 0, which is the equilibrium)
and the same velocity ẋ > 0 at t = 0 for either value of γ. Subsequently, the displacement
increases to a maximum and then decays slowly back towards the equilibrium position without
overshooting or entering into oscillation. The ticks on the time axis mark t = T0 and
t = 2 T0 , where T0 is the period that the undamped oscillator would have. This illustrates the
non-oscillatory, aperiodic nature of the motion of overdamped oscillators in general.

Notice that, all other things being equal, when the damping constant γ is increased, the maxi-
mum displacement from equilibrium is smaller and the subsequent re-approach to equilibrium
at late times is slower. This is because of the greater resistance to any motion in the presence
of a stronger damping force.
– 15 –

Fig. 3.— The motion x(t) of a critically damped oscillator (γ = 2mω0 ; thick, sharply peaked
curve) starting from rest at t = 0, compared to the behaviour of x(t) for the same physical
conditions but two stronger damping constants (γ = 4mω0 and γ = 8mω0 ; the same heavy-
damping curves shown in Figure 2). Ticks on the time axis are at t = T0 and t = 2 T0 , the
first two periods of natural oscillation for the undamped system. Critical damping allows for
the fastest possible return to equilibrium (x = 0).

Critical damping (γ = 2mω0 )

The solution obtained in Section 3 (equation [13]) for the displacement of a critically damped
oscillator is, again,

γ = 2mω0 : x(t) = e−γ t/(2m) [C1 t + C2 ] = e−ω0 t [C1 t + C2 ]

This function is drawn in Figure 3 for the oscillator with the same mass m and natural angular
frequency ω0 as in the overdamping examples in Figure 2. Here the constants C1 and C2 have
been chosen to give the system the same initial displacement x = 0 at the same initial velocity
ẋ > 0 at time t = 0 as in the overdamped examples. The overdamped x(t) curves from Figure
2 (corresponding to γ = 4mω0 and γ = 8mω0 ) are thus also reproduced in Figure 3.

This illustrates the facts that (1) like overdamping, critical damping results in non-oscillatory,
aperiodic motion, with an exponential return to equilibrium (x = 0) at late times,
but that (2) this re-approach to x = 0 is faster for γ = 2mω0 than for any larger damping
constant. In this sense, the critically damped system can be viewed as the limit of overdamped
systems. At the same time, γ = 2mω0 is the lowest value of the damping constant for which
oscillations about x = 0 can be avoided: with lower γ, the system becomes underdamped
and—as already discussed—comes to equilibrium only as the oscillation amplitude A(t) slowly
decays to zero. Thus, when all things other than γ are equal, critical damping provides for the
fastest possible return to equilibrium of an oscillator.
– 16 –

EXAMPLE

A block of mass m = 0.5 kg attached to a spring with k = 12.5 N m−1 undergoes lightly
damped oscillation with a frequency 0.2% lower than the natural frequency.

(a) What is the damping constant, γ?


The damped angular frequency is (see eq. [15] or eq. [16] above)
r
γ2
ω = ω02 −
4m2
p
where the natural (undamped) angular frequency is ω0 = k/m for a mass on a spring.
Setting ω = 0.998 ω0 in this case, squaring both sides of the equation, and solving for γ,

γ2 γ2
ω02 − = (0.998 ω0 )2 =⇒ = ω02 − (0.998 ω0 )2 =⇒
4m2 4m2
k £
γ 2 = 4m2 ω02 1 − 0.9982 = 4m2 1 − 0.9982 = 4mk 1 − 0.9982 .
£ ¤ ¤ £ ¤
m
Therefore, p
γ = 4mk [1 − 0.9982 ] = 0.316 kg s−1 .

(b) How does the amplitude vary in time?


From equation (17) above,
µ ¶ µ ¶
γt 0.316 t
A(t) = A0 exp − = A0 exp − = A0 e−0.316 t ,
2m 2 × 0.5

for t measured in seconds.

(c) How long does it take for the mechanical energy to decrease to 1% of its initial value?
As in equation (19), Etot (t) ∝ A(t)2 (including initially, E0 ∝ A20 ), so
¤2
A0 e−0.316 t = A20 e−0.632 t Etot /E0 = e−0.632 t
£
Etot ∝ =⇒

Solving this for the time when Etot /E0 = 0.01 gives

t = 7.287 sec .
p
This is approximately 5.8 oscillation periods (the undamped system has ω0 = k/m =
5 s−1 and thus T0 = 2π/ω0 ' 1.26 s; the damped system has a lower frequency and a
longer period, but only by 0.2%.)

(d) What is the critical damping constant?


By definition, γ = 2mω0 for critical damping, so in this case


r
k
γcrit = 2mω0 = 2m = 2 mk = 5 kg s−1 .
m
Lectures 9/10/11 Forced Oscillation, Resonance, Circuits

1. Energy Loss in Damped Oscillators

In our discussion of damped oscillations, we mentioned that the physical effect of the damping
force is to dissipate mechanical energy from a system (converting it into heat or some other
form of non-mechanical energy). It is straightforward to quantify the rate at which this occurs.

Recall from first-semester mechanics that the rate at which a force does work on a system is the
power, P = F v, where v is the instantaneous velocity: v = dx/dt ≡ ẋ, which is time-dependent
in general. (This follows from the definitions of power, P = dW/dt, and of the incremental
work, dW = F dx, such that dW/dt = F dx/dt = F v). In our case, with the damping force
taken to be Fdamp = −γ ẋ, this leads directly to

Pdamp = Fdamp × ẋ = (−γ ẋ) × ẋ = − γ ẋ2 (1)

Now, the total mechanical energy of an oscillator is


1 1
Etot = K + U = mẋ2 + mω02 x2 ,
2 2
(for ω0 the natural angular frequency as usual), so its rate of change is

dEtot 1 d ẋ2 1 d x2
= m + mω02
dt 2 dt 2 dt
1 d ẋ2 dẋ 1 d x2 dx
= m× × + mω02 × ×
2 dẋ dt 2 dx dt
1 dẋ 1 2 dx
= m × 2ẋ × + mω0 × 2x ×
2 dt 2 dt
2
= m ẋ ẍ + mω0 x ẋ
ẋ mẍ + mω02 x

= (2)

The chain rule has been used explicitly in the second line, and the definitions ẋ = dx/dt and
ẍ = dẋ/dt = d2 x/dt2 have been used on the second-last line.

Finally, the equation of motion for a damped oscillator,

mẍ + γ ẋ + mω02 x = 0 (3)

can evidently be re-written as


mẍ + mω02 x = − γ ẋ
so that equation (2) becomes simply
dEtot
= ẋ (−γ ẋ) = − γ ẋ2 = Pdamp (4)
dt
–2–

This shows rigorously that the rate of work done by the damping force is exactly equal to the
rate of loss of mechanical (kinetic+potential) energy from the oscillator system.

Left to itself, then, any damped oscillator would eventually lose all its mechanical energy and
come to rest at equilibrium. In nature this does not always happen, of course—sustained
oscillatory phenomena are widespread—and in practical applications we may want to prevent
this from happening. On the other hand, damping at some level can never be avoided. The
solution is to counteract damping and the dissipation of mechanical energy by applying an
external force to compensate. This motivates the study of forced or driven oscillations.

2. Forced or Driven Oscillation: The Steady State

As just discussed, if oscillations of a system are to be maintained for any significant length
of time, we require a source of energy to replace the mechanical energy dissipated by any
damping forces. Any such external, or driving force applied to this end will in general be
time-dependent (since the rate of energy loss, dEtot /dt = Pdamp = −γ ẋ2 is time-dependent
in general), and it may in principle be rather complicated. Here we consider only the case of
a harmonic driving force, which has a sinusoidal dependence on time involving a single
angular frequency:
F (t) = F0 cos(ωe t) (5)
In this formula, F0 is the amplitude of the driving force and ωe is the driving angular
frequency. That is, F (t) varies between a maximum of +F0 and a minimum of −F0 , with an
angular frequency ωe (and so an associated period of 2π/ωe ). This frequency need not bear
any particular relation at all to the natural angular frequency ω0 of the undriven oscillator,
or to any other fundamental parameter (such as the damping constant) of the system. The
subscript “e” here is used to refer exclusively to the “external” force.

Even if it may seem somewhat artificial, there are good reasons to consider only harmonic
driving forces with the simple time dependence of equation (5). First, this choice makes
it mathematically tractable to answer the basic question of how a system responds to the
external force. Second, any periodic external force can in fact be decomposed into a
sum (or superposition) of different harmonic forces with different driving frequencies
ωe . This is true no matter how complicated the full time dependence of the external force may
be in detail. Therefore, understanding the effects of a relatively simple force such as that in
equation (5) provides the groundwork for more sophisticated applications (which are, however,
beyond the scope of this module). And third, a harmonic driving force leads quickly to a very
clear illustration of the phenomenon of resonance, which is generically important.

We want to know how the displacement x(t), velocity ẋ(t), and acceleration ẍ(t) behave for an
oscillator subjected to both a damping force of the type Fdamp = −γ ẋ and a harmonic driving
force, F (t) = F0 cos(ωe t). To begin, we simply write the equation of motion in this case. The
net force acting on the system is the sum of the natural restoring force, the damping force,
and the time-varying external force:
Fnet = Frestore + Fdamp + F (t)
=⇒ mẍ = −mω02 x − γ ẋ + F0 cos(ωe t)
–3–

or, in a more standard form,

mẍ + γ ẋ + mω02 x = F0 cos(ωe t)


γ F0
or ẍ + ẋ + ω02 x = cos(ωe t) (6)
m m

As might easily be imagined, the solution x(t) of this equation correponds to motion that is
somewhat more complicated than what we have dealt with so far. Nevertheless, we can make
headway with the problem by thinking about what we expect the system to do first at early
times and then at late times:

(1) If we think of the external force driving force as being “switched on” at some time t = 0,
then the equation of motion for times prior to this would be nothing more than equation (3)
for an undriven (or “free”) damped oscillator, since this is just equation (6) with the right-
hand side set to zero. We already know the solutions for the undriven damped oscillator:
depending on the value of γ vs 2mω0 , x(t) will be one of the underdamped, critically damped,
or overdamped solutions discussed in Lectures 6–8. This motion will persist for a while after the
external force has been “switched on,” and it will in fact dominate the motion of the system at
very early times t > 0. Eventually, however, this familiar, under- or over- or critically damped
motion will die away, dissipated by the damping force. It is therefore not important in the
long term. The early stage when it is important is called the transient stage, and x(t) in this
stage is called the transient solution.

(2) As time goes on, the full x(t) becomes first a complicated superposition of the transient
solution and an “extra” motion that comes directly in response to the external force. Ulti-
mately, when the transient solution has decayed essentially completely, the only motion left is
the response to the driving force. In this limit x(t) should be governed by the characteristics of
the external F (t), i.e., by F0 and ωe in particular. This late-time situation is referred to as the
steady state, and the displacement x(t) in this stage is called the steady-state solution.

In mathematical terms, the general solution x(t) to the differential equation of motion for a
forced oscillator is the sum of the solution xh (t) to the associated homogeneous equation (3)
plus a particular solution xp (t) to the full, non-homogeneous equation (6). The solution to
the homogeneous equation is the transient that dies away relatively quickly, leaving only the
particular solution xp (t) in the steady state at late times. We do not discuss the transient
stage of the forced oscillator in any more detail; instead, we focus entirely now on the steady-
state solution xp (t) and the long-term behaviour. However, see Appendix B for illustrations
of general solutions including both the transient and the steady-state contributions to the full
x(t).

To find the steady-state solution xp(t), we first reason that it should involve oscillation
at the same angular frequency as the driving force, with some amplitude to be
determined and possibly with a phase lag between the variation of the external force
and any response in the displacement. That is, we hypothesize that

xp (t) = A cos(ωe t − δ) (steady state) (7)

where A is the oscillation amplitude and δ is the phase lag (also sometimes called the phase
angle; measured in radians).
–4–

Given the hypothetical form (7) for the steady-state solution, we need to confirm that it does
in fact satisfy the equation of motion (6). In the course of doing so, we can infer self-consistent
constraints on the values of A and δ. It is in fact easiest to do this by considering the related
complex differential equation,
γ F0 iωe t
z̈p + żp + ω02 zp = e (8)
m m
in which the complex function zp (t) is assumed to be of the form

zp (t) = A ei(ωe t−δ) (9)

with A and δ both real constants.

The connection of equations (8) and (9) with the physical problem at hand is that the driving
force we have assumed, F0 cos(ωe t), is precisely the real part of (F0 /m)eiωe t in equation (8);
while the steady-state solution xp (t) = A cos(ωe t − δ) that we seek is just the real part of
the complex function zp (t) = Aei(ωe t−δ) in equation (9). Thus, if we can find A and δ to
make equation (9) a valid solution of the differential equation (8), then the same A and δ are
guaranteed to make the real xp (t) in equation (7) a solution to the original equation (6). The
advantage to working with complex numbers and functions in this case is that the algebra
involved in differentiating complex exponentials is easier than—but equivalent to—the algebra
involved in differentiating sine and cosine functions explicitly. [See Appendix A for details of
how to solve for xp (t) without using complex numbers.]

Thus, we now calculate the first two time derivatives of the complex function zp (t):

zp (t) = A ei(ωe t−δ) =⇒ żp (t) = iωe A ei(ωe t−δ)


(10)
z̈p (t) = (iωe )2 A ei(ωe t−δ) = −ωe2 A ei(ωe t−δ)

Putting these expressions directly into equation (8) then leads to

γ F0 iωe t
−ωe2 A ei(ωe t−δ) + × iωe A ei(ωe t−δ) + ω02 × A ei(ωe t−δ) = e
m m
Next, multiplying through by eiδ and then cancelling factors of eiωe t from both sides gives
γ F0 iωe t
−ωe2 A eiωe t + iωe A eiωe t + ω02 A eiωe t = e × eiδ
m m
γ F0 iδ
=⇒ −ωe2 A + iωe A + ω02 A = e
m m
Finally, we invoke the definition eiδ = (cos δ + i sin δ):

γ F0
−ωe2 A + iωe A + ω02 A = (cos δ + i sin δ)
m m
γ F0 F0
=⇒ (ω02 − ωe2 ) A + i ωe A = cos δ + i sin δ (11)
m m m

For equation (11) to be true, the real and imaginary parts of it must be separately satisfied.
That is, the terms not mulitplied by i on the left- and right-hand sides must equal each other,
and the terms that are multiplied by i on the left- and right-hand sides must equal each other.
–5–

We have therefore obtained a pair of simultaneous equations for the amplitude A and phase
lag δ of the steady-state oscillation of a forced oscillator:
F0
ω02 − ωe2 A =

cos δ (12)
m
γ F0
and ωe A = sin δ . (13)
m m
To solve these equations for the two constants:

• First, square both sides of equation (12) and square both sides of equation (13), and then
add the left- and right-hand sides of the two squared equations:
γ  2  2
 2 2
 2 2 F0 F0
ω0 − ω e A + ωe A = cos δ + sin δ ,
m m m
which simplifies to

γ 2 ωe2 F02  2 F02


 
2 2 2 2 2
 
A ω0 − ω e + = cos δ + sin δ =
m2 m2 m2

(since sin2 δ + cos2 δ = 1 for any angle δ). Isolating A and taking square roots therefore
gives
F0 /m
A = q 2 (14)
2 2 2
ω0 − ωe + γ ωe /m 2 2

for the displacement amplitude of a driven oscillator in the steady state.

• Second, divide equation (13) by equation (12):

(γ/m) ωe A (F0 /m) sin δ


2 2
= ,
(ω0 − ωe )A (F0 /m) cos δ

which gives
γωe /m
tan δ = (15)
ω02 − ωe2
for the phase lag between force and displacement in the steady state.

It is notable that there is no freedom left in xp (t) = A cos(ωe t − δ) once the amplitude F0 and
the angular frequency ωe of the external, driving force are specified. Together with the natural
angular frequency ω0 , the damping constant γ, and the mass m of the oscillator, these things
specify completely the value of the amplitude and the phase angle in the long-term, steady
state behaviour of the system after the initial transient stage has died away. This is rather
unlike the situation for simple harmonic oscillation, or even damped but undriven oscillation,
where there are always two independent constants that reflect initial conditions. Such free
constants are absent in the present context because in the steady state a forced oscillator
responds directly to the external force and has “forgotten” its initial conditions.

Figures 1 and 2 illustrate the steady-state displacement amplitude and phase lag just derived
(equations [14] and [15]), as functions of the driving angular frequency ωe applied to systems
with various relative strengths of the damping constant γ.
–6–

Fig. 1.— Amplitude A of the steady-state displacement of a driven oscillator, as a function of


the angular frequency ωe of the external force (equation [14]). Broader curves correspond to
larger values of the damping constant γ.

Fig. 2.— Phase difference, or phase lag, between the external force F = F0 cos(ωe t) and the
steady-state displacement xp (t) = A cos(ωe t − δ) of a driven oscillator (equation [15]). Going
from bottom to top on the left-hand side of the plot, or from top to bottom on the right-hand
side, the sequence of curves corresponds to increasing values of the damping constant, γ.
–7–

The shapes of the curves for the displacement amplitude A in Figure 1 are typical of the
response curves for forced oscillators. These are discussed in more detail in Section 3 below.
First, it is worth noting some general features of the steady-state motion, with reference both
to Figures 1 and 2 and to equations (14) and (15).

Looking at the lower left-hand corner of Figure 1, in the limit that the driving frequency
ωe −→ 0 [which corresponds to a constant external force, F0 cos(ωe t) −→ F0 cos(0) = F0 ],
the displacement amplitude tends to a constant value, independent of the damping constant:
setting ωe = 0 in equation (14) yields A −→ F0 /(mω02 ). At the same time, in Figure 2 (or
equation [15]) the phase lag δ tends to zero as ωe goes to zero, meaning that the displacement
as a function of time tends to xp (t) = A cos(ωe t − δ) −→ A cos(0 − 0) = A = F0 /(mω02 ).
Therefore (and as is also discussed further below), in this limit of low-frequency driving the
external force works to cancel the natural restoring force (Frestore = −mω02 ) and simply
shifts the oscillator to a new, non-zero equilibrium position (at xp = A) where the system rests
with no oscillation in the steady state.

In the opposite limit of very large driving frequencies, ωe ≫ ω0 or ωe −→ ∞, the steady-state


displacement amplitude becomes very small (A −→ 0, as can be seen in the far right-hand
corner of Figure 1 or directly from equation [14]), while the phase lag tends to δ −→ π radians
(see Figure 2), so the displacement becomes 180◦ , or one-half of a cycle, out of phase with the
driving force. As such (and again as discussed below), in this limit of high-frequency driving
the external force works to replace the natural restoring force of the oscillator. The result
is SHM at the angular frequency of the applied force but with a very small amplitude in the
steady state.

In between these extremes, for values of the driving angular frequency ωe at or near
the natural angular frequency ω0 , the displacement amplitude can be very large. This is
one manifestation of the phenomenon of resonance, which is discussed further in Section 3.
Another aspect of this is that the phase lag δ is approximately π/2 rad = 90◦ , which—as we
shall see in Section 3—means that in the steady state the instantaneous velocity, ẋp (t), is nearly
in phase with the driving force. In fact, when ωe = ω0 exactly, the steady-state velocity
is exactly in phase with the driving force, which therefore cancels the damping
force exactly. The response of the system to the external force is therefore greatest
for driving frequencies comparable to the natural angular frequency, and as a result
the displacement amplitude is a maximum when the driving angular frequency is near ωe ≃ ω0
in general (although, strictly speaking, A is maximized for an ωe,max that is less than ω0 ).

The strength of the peak in A as a function of ωe depends on the amount of damping in the
system. Very lightly damped systems, with low values of the constant γ, have the steady-state
A peaking very strongly and very close to ωe = ω0 . More heavily damped systems (higher γ)
are more efficient at dissipating the energy input by the driving force and thus have steady-
state displacements that do not respond as dramatically to the external driving and are not
as sharply “tuned” to ωe = ω0 . Hence the curves for A vs ωe in Figure 1 are broader for

larger values of the constant γ. In fact, if the damping is so strong that γ ≥ 2 mω0 ,
then the steady-state displacement amplitude shows no resonance-type maximum
at all. Instead, in this case A decreases monotonically as the driving frequency increases from
ωe = 0, through ωe = ω0 , and on to ωe −→ ∞.
–8–

3. Low-Frequency Driving, High-Frequency Driving, and Resonance

In this section we expand on the qualitative discussion, at the end of Section 2, of low-frequency
driving, high-frequency driving, and resonance (driving at or near the natural angular fre-
quency) in a forced oscillator. Throughout, we will be referring to our solution for the steady-
state displacement xp (t) of a driven oscillator and the associated velocity ẋp (t) and acceleration
ẍp (t) as functions of time. From Section 2, then,

external F (t) = F0 cos(ωe t) =⇒ xp (t) = A cos(ωe t − δ)


ẋp (t) = −ωe A sin(ωe t − δ) (16)
ẍp (t) = −ωe2 A cos(ωe t − δ)

in which the amplitude A and the phase lag δ are given by (from equations [14] and [15])
F0 /m γωe /m
A = q and tan δ = (17)
2 ω02 − ωe2
ω02 − ωe2 + γ 2 ωe2 /m2

There are three situations of particular physical interest here:

• Low-frequency driving: ωe ≪ ω0
This is when the external force varies much more slowly than the natural restoring force,
Frestore = −mω02 x. In this limit, the driving angular frequency ωe can be approximately
ignored relative to the natural angular frequency ω0 , so the denominator in the expression
for tan δ in equation (17) is (ω02 −ωe2 ) ≈ ω02 . Thus, tan δ ≈ [(γωe )/m]/ω02 ; and since ωe /ω02
is a very small number, tan δ ≈ 0. That is, the steady-state displacement is nearly
in phase with the external force. Moreover, the denominator in the expression
for the steady-state displacement amplitude A in equation (17) becomes approximately
ω04 + γ 2 ωe2 /m2 ≈ ω02 (because ω02 ≫ ωe2 and ω04 ≫ ωe2 under the square root). Equations
p

(16) then imply


F0 F0
xp (t) ≃ cos(ωe t) −→
mω02 mω02
ωe F0
ẋp (t) ≃ − sin(ωe t) −→ 0
mω02
ω 2 F0
ẍp (t) ≃ − e 2 cos(ωe t) −→ 0
mω0
where the limit at the end of each line corresponds to letting the driving frequency ωe go
to zero (giving a time-independent external force of amplitude F0 ). This shows directly
that in the steady state of a forced oscillator driven at low frequencies, the external force
simply shifts the system to a new equilibrium (xp = F0 /mω02 ), where it sits with zero
velocity and acceleration. In other words, a driving force with a very low angular
frequency effectively cancels the natural restoring force.

• High-frequency driving: ωe ≫ ω0
This is when the external force varies much more rapidly than the natural restoring
force, Frestore = −mω02 x. In this limit, the natural angular frequency ω0 can be ap-
proximately ignored relative to the driving angular frequency ωe , so the denominator
–9–

in the expression for tan δ in equation (17) is approximately (ω02 − ωe2 ) ≈ −ωe2 . Thus,
tan δ ≈ −[(γωe )/m]/ωe2 = −γ/(mωe ). In the extreme limit of ωe −→ ∞, tan δ approaches
0 but is negative. Thus, δ tends to π radians (180◦ , or half a cycle) in this case. That is,
the steady-state displacement is nearly in anti-phase with the external force.
Moreover, the denominator in the expression for the steady-state displacement amplitude
A in equation (17) becomes approximately ωe4 + γ 2 ωe2 /m2 ≈ ωe2 (because ωe2 ≫ ω02 and
p

ωe4 ≫ ωe2 under the square root). Equations (16) then imply

F0 F0
xp (t) ≃ cos(ωe t − π) = − cos(ωe t)
mωe2 mωe2
F0
ẋp (t) ≃ + sin(ωe t)
mωe
F0
ẍp (t) ≃ + cos(ωe t)
m
The final equation, for the steady-state amplitude, shows directly that

mẍp = F0 cos(ωe t) = F (t)

as if the external force were the only one acting on the system; while the result for xp (t)
itself shows that ẍp = −ωe2 xp , so the motion in the steady state is “regular” simple
harmonic oscillation at the angular frequency, ωe , of the external force (in fact, this is
the case for any value of ωe ). As such, a driving force with a very high angular
frequency effectively replaces the natural restoring force. The amplitude of
the steady-state oscillation is relatively small, however: in the limit ωe −→ ∞, the
maximum possible displacement F0 /(mωe2 ) −→ 0 and the maximum possible velocity
F0 /(mωe ) −→ 0 (even though the maximum acceleration is a non-zero constant, F0 /m).

• Driving at the natural frequency: ωe = ω0


This case is of interest both because of the implied “tuning” to the natural state of
the oscillator being driven, and because the results for the steady-state displacement
amplitude and phase lag become very simple. Simply putting ωe = ω0 into the formulae
for A and tan δ in equation (17) leads to

F0 π
A = and tan δ −→ ∞ =⇒ δ =
γω0 2

Thus, equations (16) become

F0 F0
xp (t) ≃ cos(ω0 t − π/2) = sin(ω0 t)
γω0 γω0
F0
ẋp (t) ≃ cos(ω0 t)
γ
ω0 F0
ẍp (t) ≃ − sin(ω0 t)
γ
The second of these equations is particularly important: it shows that

F0 cos(ω0 t) = γ ẋp (t)


– 10 –

which is to be compared to Fdamp = −γ ẋp for the damping force in the steady state, and
to F (t) = F0 cos(ω0 t) for the external force (given that ωe = ω0 ). That is,

ωe = ω 0 =⇒ F (t) = − Fdamp

so the driving force cancels the damping force exactly in this case. This can-
cellation is also implied by the fact that δ = π/2 rad, or 90◦ . This means that the
steady-state displacement is one-quarter cycle out of phase with the external force, and
therefore that the steady-state velocity is precisely in phase with the external force (recall
that velocity and displacement are normally π/2 or 90◦ out of phase with each other in
simple harmonic oscillation).
The maximum possible displacement in the steady state, when ωe = ω0 , is the collection
of constants multiplying sin(ω0 t) in the expression for xp (t) just above:
 
F0 mω0 F0
|xp |max = =
γ ω0 γ mω02
which is evidently larger, by a factor mω0 /γ, than the maximum displacement F0 /(mω02 )
achieved when the driving force cancels the restoring force (low-frequency driving). Sim-
ilarly, the maximum possible acceleration when ωe = ω0 is
 
ω0 F0 mω0 F0
|ẍp |max = =
γ γ m
which is larger, by the same factor mω0 /γ, than the maximum acceleration F0 /m when
the driving force replaces the restoring force (high-frequency driving).
This factor of mω0 /γ is known as the Q-value:
mω0
Q ≡
γ
If the damping constant γ is very low, then Q can be very large, and driving at the
natural angular frequency can strongly amplify the steady-state motion that would ob-
tain for driving forces at much lower or higher frequencies. Notice, however, that the
“amplification” is Q > 1 only if the damping constant is low enough that γ < mω0 .
This implies that the forcing of lightly damped oscillators is of most interest
(recall that, in the absence of an external driving force, an oscillator is underdamped as
long as γ < 2mω0 ). In fact, in the limit of no damping at all (γ −→ 0, although this
can never actually be achieved in a real system), we would have Q −→ ∞ and so the
amplitudes of the displacement, velocity, and acceleration in the steady state (which are
all proportional to Q) would also tend to infinity.

The above leads to the basic conclusion that interaction between a system and an ex-
ternal force is maximized for a driving frequency at or near the natural angular
frequency, i.e., for ωe ≃ ω0 . This is the essence of resonance. Resonance is allowed by the
cancellation of the damping force by the driving force when ωe = ω0 (or a near-cancellation
for ωe ≃ ω0 ); and it is characterized by the amplification of steady-state amplitudes (Q > 1)
around ωe ≃ ω0 . Figure 3 illustrates this by showing response curves, which are the ampli-
tudes of the steady-state xp (t), ẋp (t), and ẍp (t) as functions of ωe . From equations (16), these
amplitudes are A, ωe A, and ωe2 A, with A itself given in equation (17).
– 11 –

Fig. 3.— Response curves, showing the dependence of the steady-state displacement amplitude
|xp |max = A, the steady-state velocity amplitude |ẋp |max = ωe A, and the steady-state acceler-
ation amplitude |ẍp |max = ωe2 A, all as functions of the driving angular frequency ωe . Different
curves in any one panel correspond to different damping constants, γ; the narrower curves are
always for lower γ. For γ ≥ 2mω0 there is no “resonance peak” in either the displacement
amplitude or the acceleration amplitude, for any value of ωe .
– 12 –

Displacement Amplitude The plot of displacement amplitude, A vs ωe in the top panel


of Figure 3, is similar to Figure 1. The different curves correspond to different values of the
damping constant γ, for fixed values of oscillator mass m, natural angular frequency ω0 , and
external force amplitude F0 . Narrower curves, with sharper and higher peaks in the amplitude
for ωe near ω0 , correspond to lower damping constants.

In general, the displacement amplitude is, again,


F0 /m
A = q 2
ω02 − ωe2 + γ 2 ωe2 /m2

Differentiating this with respect to ωe and setting the result equal to zero yields the driving
angular frequency that maximizes A. This defines the condition of amplitude resonance:
dA
q
= 0 =⇒ ωe,max = ω02 − γ 2 /(2 m2 ) < ω0
dωe
This result is proved in Appendix C below. As indicated, it shows that amplitude resonance
occurs at driving angular frequencies ωe,max less than the natural ω0 in general. If the
damping constant γ is not too large, then the angular frequency for amplitude resonance is not
too much less than ω0 . However, if there is to be amplitude resonance at all, then ωe,max must
be a real number, which means the quantity under the square root must be positive. That is,

ω02 > γ 2 /(2m2 ) =⇒ γ < 2 mω0

is required to allow amplitude resonance. The lowest curve in the top panel of Figure 3, which

has no peak, corresponds to γ = 2 mω0 . Recall that an undriven oscillator is underdamped if
γ < 2mω0 . Thus, only underdamped oscillators can display amplitude resonance when driven;

but more than this, only sufficiently underdamped systems, which have γ < 2 mω0 <
2mω0 are susceptible to the phenomenon.

Velocity Amplitude The plot of steady-state velocity amplitude, ωe A vs ωe in the middle


panel of Figure 3, shows the response curves for several values of the damping constant, with
lower γ yielding a sharper peak. The curves differ from those for A vs ωe in that the velocity
amplitude goes to zero for both ωe −→ 0 (constant external force; low-frequency driving) and
ωe −→ ∞ (high-frequency driving).

The velocity amplitude is (from, e.g., equation [16])


ωe F0 /m
ωe A = q 2
ω02 − ωe2 + γ 2 ωe2 /m2

Differentiating this with respect to ωe and setting the result equal to zero yields the driv-
ing angular frequency for which the velocity amplitude is maximized. This defines velocity
resonance:
d (ωe A)
= 0 =⇒ ωe = ω0
dωe
which is also proved in Appendix C below. This can occur for any value of the damping
constant, although as γ increases the resonance peak in the amplitude ωe A becomes very
weak and the phenomenon is of little physical importance.
– 13 –

Acceleration Amplitude The plot of steady-state acceleration amplitude, ωe2 A vs ωe


in the bottom panel of Figure 3, shows curves for several values of the damping constant,
with lower damping yielding a sharper peak. The acceleration amplitude tends to zero in the
low-frequency driving limit, ωe ≪ ω0 , and to a constant F0 /m independent of the damping
constant γ in the limit of high-frequency driving, ωe ≫ ω0 .

The acceleration amplitude is (from equation [16] again)


ωe2 F0 /m
ωe2 A = q 2
ω02 − ωe2 + γ 2 ωe2 /m2

Differentiating this with respect to ωe and setting the result equal to zero yields the driving
angular frequency for which the acceleration amplitude is maximized. This is
d (ωe2 A) .q
= 0 =⇒ ωe = ω02 ω02 − γ 2 /(2 m2 ) > ω0
dωe
Again, this is proved in Appendix C. As indicated, the driving frequency that maximizes the
steady-state acceleration amplitude is greater than the natural angular frequency ω0 in
general—although it will be near ω0 if γ is not too large. As with the displacement amplitude,

a peak in the acceleration amplitude can only occur at all provided that γ < 2 mω0 (so the
square root just above is a real number), and therefore this is relevant only for sufficiently
underdamped oscillators.

It is worth emphasizing again that all of the response curves of Figure 3 are broader for higher
values of the damping (to be precise, for higher γ/m at a given ω0 ), meaning that there is
a larger range of driving frequency ωe over which the steady-state response of the system
(whether considered in terms of displacement, velocity, or acceleration amplitude) is close to
the resonance maximum. For this reason, the ratio γ/m is called the width. This has units
of (kg s−1 )/kg = s−1 .

4. Power in Forced Oscillators

We motivated our discussion of forced oscillation by noting, in Section 1 above, that a damping
force of the type Fdamp = −γ ẋ works continually to remove mechanical energy from an oscilla-
tor, at the rate Pdamp = dEtot /dt = −γ ẋ2 . We return to this point now, to see quantitatively
how the application of an external harmonic force affects the situation in the steady state.

First, because the result Pdamp = −γ ẋ2 is completely general (being based only on the definition
of power and the assumed form of the damping force—not on any other details of any specific
system), we have, in the steady state of a driven oscillator,

Pdamp = − γ ẋ2p = − γ [−ωe A sin(ωe t − δ)]2 = − γωe2 A2 sin2 (ωe t − δ)

where the general form of the steady-state velocity, ẋp (t), can be seen in equation (16). A and
δ are the usual displacement amplitude and phase lag, as in equation (17).

Evidently, this rate of work done by the damping force is itself a function of time, varying
along with the instantaneous steady-state velocity, ẋp (t). However, even without knowing the
– 14 –

details at every instant of time we can still say what the average damping power is over any
full oscillation cycle (or indeed any whole number of cycles). This is because one oscillation
cycle corresponds to a change of exactly 2π in the phase (ωe t − δ) of the steady-state velocity
(and also in the phase of the displacement and acceleration). But then, it is a standard result
that the average value of sin2 θ for any angular variable θ over the range 0 ≤ θ ≤ 2π
is just 1/2. (This is proved in Appendix D below, though it is only the result itself that we
need here.) Given this, if we use angular brackets hi to denote the average of a quantity over
one oscillation cycle, we have
1
hPdamp i = − γωe2 A2 × sin2 (ωe t − δ) = − γω 2 A2
2 e
for the average rate of energy dissipation by the damping force in the steady state.
Notice that hPdamp i is still negative, since the damping force by itself still leads to a net loss
of energy. The value of hPdamp i can always be evaluated for any given values of ωe , F0 , ω0 , m,
and γ, by using equation (17) for the displacement amplitude A.

In addition to the (negative) work done by the damping force, work is also done by the
external driving force. The rate at which this happens is (again, by the definition of power)
Pext = F (t) × ẋp (t). Putting F (t) = F0 cos(ωe t) and ẋp (t) = −ωe A sin(ωe t − δ) into this,

Pext = F0 cos(ωe t) × [−ωe A sin(ωe t − δ)]


= F0 cos(ωe t) × [−ωe A sin(ωe t) cos δ + ωe A cos(ωe t) sin δ]
= −ωe AF0 cos δ sin(ωe t) cos(ωe t) + ωe AF0 sin δ cos2 (ωe t)

where the second line has made use of the fact that sin(Θ − Φ) = sin Θ cos Φ − cos Θ sin Φ for
any two angles Θ and Φ. Once again, this rate of work is time-varying, with a dependence
that is even more complicated than that of Pdamp above. However, the result is simplified
considerably when we average again over one full oscillation cycle (or a whole number of
cycles). First, the cycle-average hcos2 (ωe t)i is 1/2 (just as the average of sin2 is); and second,
the average hsin(ωe t) cos(ωe t)i is just 0 (this is also proved in Appendix D). Therefore,

hPext i = −ωe A F0 cos δ × hsin(ωe t) cos(ωe t)i + ωe A F0 sin δ × cos2 (ωe t)


1
= 0 + ωe A F0 sin δ ×
2
1
= ωe A F0 sin δ
2
But then, in our derivation of the amplitude A and phase lag δ for the steady state of a forced
oscillator, we found that (see equation [13] in Section 2)
γ F0
ωe A = sin δ =⇒ F0 sin δ = γ ωe A
m m
and so we finally have
1
hPext i = γω 2 A2 = − hPdamp i
2 e
for the power supplied to the system in the steady state.

This shows directly that, in the steady state, the average power input to (or absorbed
by) a driven oscillator from the external force exactly balances the average rate
– 15 –

of energy dissipation by the damping force. The average rate of change of mechanical
(kinetic+potential) energy is therefore
 
dEtot
= hPdamp i + hPext i = 0
dt

This is as expected. Such conservation of energy is what allows oscillations of the system to
be sustained for any appreciable length of time.

Notice that the product ωe2 A2 in the expression for either hPdamp i or hPext i is simply the
square of the velocity amplitude in the steady state (see equation [16] again). The velocity
amplitude is maximized in velocity resonance, which is when the driving angular frequency
equals the natural angular frequency. Thus, the condition of velocity resonance, ωe = ω0 ,
also maximizes both the power absorbed by the forced oscillator and the power
dissipated by the damping force. Figure 4 shows the response curve for the average power
(either hPext i or −hPdamp i). This is essentially proportional to the square of the response curve
for the steady-state velocity amplitude in the middle panel of Figure 3. As such, the power
peaks clearly at ωe = ω0 for any value of the damping constant γ. The height and width of
the peak themselves depend on γ, qualitatively as indicated on the plot.

Finally, our standard formula for A (e.g., in equation [17]) allows an easy calculation of the

Fig. 4.— Response curve for the (cycle-averaged) power in the steady state of an oscillator
driven by a harmonic external force, for a few different values of the damping constant γ.
The peaks are maximum hPext imax = F02 /2γ and hPdamp imax = −F02 /2γ. The maximum is
achieved with adriving ωe = ω0 , which is also the condition for velocity resonance.
– 16 –

maximum power when ωe = ω0 (the heights of the peaks of the curves in Figure 4):

F0 1 F2
ωe = ω 0 =⇒ A = =⇒ hPext imax = −hPdamp imax = γ (ωe A)2max = 0
γω0 2 2γ

For a very lightly damped system (small γ) being driven at ωe = ω0 , the average power
absorbed from the driving force can thus become very large. Moreover, all of this supplied
power is ultimately dissipated by the damping force, essentially being put into the system as
heat. This is an example of why a maximum interaction between system and environment,
which is what resonance provides, has the potential to be problematic in some respects even if
it may be appealing for other reasons.

5. Application: Oscillation and Resonance in Electrical Circuits


Further discussion of some aspects of circuits relevant to this section can be found in
Chapters 32.4, 32.5, 33.6, 33.7, and 33.8 of Benson, University Physics

Any number of examples of forced mechanical oscillations and resonance could be given to illus-
trate “real-world” applications of the ideas and formulae developed above. Familiar ones may
be a child pumping a swing; bridges swaying or collapsing; or a soprano shattering a glass with
a well-pitched note. While instructive, all of these examples depend to some extent on idealiza-
tions to model complicated physical systems in terms of the relatively simple, linear damping
forces and harmonic external forces that we have assumed throughout our analysis. In many
ways a much cleaner example, which comes very close in reality to matching the mathematical
assumptions we have made, is of oscillations and resonance in electrical circuits. This
is also of interest because, even though it involves non-mechanical oscillations (of charge,
current, and voltage), rather than the explicitly mechanical oscillations (of displacement, ve-
locity, and acceleration) that we have focussed on to this point, the mathematical results that
we have obtained nevertheless carry over directly to the new setting.

One possible element in a circuit is an inductor (or “coil”): a device in which a passing current
causes a changing magnetic field, which in turn induces a voltage (or “electromotive force,”
EMF) across the inductor. This induced voltage is

dI
VL = L (18)
dt
where L is a constant called the inductance and I is the current, which is time-dependent in
general.

Another circuit element is a capacitor, which stores charge and has an associated voltage of
Q
VC = (19)
C
where Q is the stored charge, and the constant C is the capacitance.

A third element is a resistor, which dissipates current. The voltage across a resistor is

VR = IR (20)
– 17 –

where the constant R is the resistance and I is again the current.

The final element that we allow for is a possible source of external voltage, which in general
may vary in time:
source voltage: VS (t) (21)

Figure 5 shows a simple circuit in which the four elements just mentiones are connected in
series. We wish now to describe the flow of charge (i.e., current) through this circuit.

The sum of voltages across the three RLC elements illustrated must equal the voltage supplied
by the source element. That is, bringing in all of equations (18)–(21), we require

VL + VR + VC = VS (t)
dI Q
=⇒ L + RI + = VS (t) (22)
dt C
But then we have that current is the rate of change of charge:
dQ
I =
dt
so we can also write equation (22) as

d2 Q dQ 1
L 2
+ R + Q = VS (t)
dt dt C
d2 Q R dQ 1 VS (t)
or 2
+ + Q = (23)
dt L dt LC L

The connection with the forced mechanical oscillations that we have been discussing now
becomes clear, since equation (23) for the charge Q(t) is identical in form to our equation

Fig. 5.— Series circuit consisting of a resistor R, an inductor L, a capacitor C, and an


“external” time-dependent voltage source, VS (t).
– 18 –

of motion for an oscillator driven by an external force, F (t):


γ F (t)
ẍ + ẋ + ω02 x =
m m
To make the connection explicit and precise, we need only to adopt a simple scheme of analogies
between the RLC series circuit and our usual mechanical oscillator:
Q ←→ displacement x
I = dQ/dt ←→ velocity ẋ
L ←→ mass m
R ←→ damping constant γ

1/ LC ←→ natural angular f requency ω0
VS (t) ←→ external f orce F (t)

With nothing more than a change of variables, we may then infer the basic behaviour of series
RLC circuits in a few different situations.

• VS (t) ≡ 0 and R = 0 : Simple Harmonic Oscillation


If there is no “external” voltage and no resistor, leaving only an inductor and a capacitor,
then the “equation of motion” (23) for the circuit becomes simply

d2 Q 1 d2 Q 1
2
+ Q = 0 =⇒ 2
= − Q
dt LC dt LC
Hence the charge undergoes simple harmonic oscillation: Q(t) = Q0 sin(ω0 t + φ0 ), for
some constants Q0 and φ0 (determined by initial conditions) and a natural angular

frequency ω0 = 1/ LC . The current, I(t) = dQ/dt, oscillates similarly (though 90◦ out
of phase with the charge). See Chapter 32.4 of Benson for further discussion.

• VS (t) ≡ 0 and R 6= 0 : Damped Oscillation


A resistor added in series to an inductor and a capacitor leads to damping of the
charge/current, and the “equation of motion” (23) reflects this:

d2 Q R dQ 1
2
+ + Q = 0
dt L dt LC
The solution for Q(t) will be formally the same as one of the solutions for the displace-
ment x(t) of underdamped, critically damped, or overdamped mechanical oscillators from
Lectures 6–8; a substitution of symbols just has to be carried out, following the scheme
above. The time-dependent current then follows as I(t) = dQ/dt. Whether the damping
is light, critical, or heavy depends on whether the resistance R (the analogue of the damp-
p
ing constant γ) is less than, equal to, or greater than the value of 2 L/C (the analogue
of 2mω0 in the mechanical case). See Chapter 32.5 of Benson for further discussion.

• VS (t) = V0 cos(ωe t) and R 6= 0 : Forced or Driven Oscillation (AC Circuit)


A sinusoidally varying voltage source is the essence of an alternating-current (AC) circuit.
With such a source, the “equation of motion” (23) is

d2 Q R dQ 1 V0
2
+ + Q = cos(ωe t)
dt L dt LC L
– 19 –

which is directly comparable to equation (6) for the harmonic driving of a mechanical
oscillator. In this case, the voltage amplitude V0 takes the place of the force amplitude
F0 . We have given the driving voltage frequency the same symbol, ωe , as the driving
frequency of the external mechanical force in our earlier analysis. The natural angular

frequency of the circuit is still ω0 = 1/ LC , and all of the other electrical ↔ mechanical
replacements listed above still apply.
Our entire discussion about the solution to the forced-oscillator equation of motion car-
ries over to the series AC circuit. In particular, the time-dependent charge Q(t) has a
transient component that is damped (by the resistor) and dies away relatively early on.
After this a steady-state
Q(t) = Q0 cos(ωe t − δ)
dominates, in which the amplitude Q0 is analogous to the displacement amplitude A,
and the phase lag δ between the charge and the source voltage is the analogue of the lag
between steady-state displacement and an external force. Thus, simply by changing the
variables in equation (17),
V0 /L Rωe /L
Q0 = q and tan δ = 1 2
LC − ωe
1
2
LC − ωe2 + R2 ωe2 /L2

In the context of circuits, we are generally more interested in the current (the analogue
of “velocity”) than in the charge (“displacement”) per se. In the steady state, given
VS (t) = V0 cos(ωe t) and with Q(t) as just above, the current is
dQ
I(t) = = − ωe Q0 sin(ωe t − δ)
dt
so the current amplitude is
ωe V0 /L
I0 = ωe Q0 = q 2
1
LC − ωe2 + R2 ωe2 /L2

After some algebra to bring a factor of ωe /L out of the denominator to cancel with the
ωe /L in the numerator, this can be re-written as
V0 V0
I0 = r ≡
1
2 Z
ωe C − ωe L + R 2

where the impedance Z, defined as


s 2
1
Z ≡ − ωe L + R 2
ωe C
is derived by other means in Chapter 33.6 of Benson.
The average power input to the circuit by the voltage source—and dissipated by the
resistor—follows from replacing the damping constant γ by R and the velocity amplitude
ωe A by the current amplitude I0 in our previous expression from Section 4:
1 1 1 V02 R
hPext i = − hPdamp i = γω 2 A2 ←→ hP i = (R)(I02 ) =
2 e 2 2 Z2
– 20 –

This result for the circuit is is derived by other means in Chapter 33.8 of Benson.
Finally, the phenomenon of resonance also carries over to the AC circuit (see also Section
33.7 of Benson). Again, we are interested primarily in the current, and since this is
the analogue of velocity we know that the resonance of interest occurs when the
driving voltage frequency matches the natural angular frequency exactly:

ωe = ω0 = 1/ LC . In this resonance, the current is exactly in phase with the
driving voltage; the impedance Z just above reduces simply to Z = R; the current
amplitude is I0 = V0 /R; and the average power absorbed (and dissipated) by the
circuit is just hP i = (1/2)I02 R = (1/2)V02 /R.
The response curve for the steady-state current in an AC circuit looks like the middle
panel of Figure 3 in Section 3 (the response curve for the velocity amplitude of a me-
chanical oscillator), while the response curve of the power looks like Figure 4 in Section

4. In both cases, the curves peak at ωe = ω0 = 1/ LC, and the peak is higher and
narrower for smaller values of the resistance R.
The resonance phenomenon and the shape of the current/power response curves are di-
rectly of importance for the tuning of radios and televisions: When a broadcast signal
(electromagnetic wave) with a particular angular frequency ωe reaches the circuit, ad-

justing the circuit elements so that the natural angular frequency ω0 = 1/ LC matches
the broadcast ωe well (i.e., for resonance) allows the signal to be received. If the response
curve for the circuit is narrow enough, then reception will fall off rapidly for frequencies
even slightly away from the resonance, preventing interference from other broadcasts.
– 21 –

APPENDICES
The material below is meant to aid understanding but is not examinable.

A. Solving for the Steady-State x(t) Without Complex Numbers

In Section 2, we solved the equation of motion for a forced oscillator in the steady state by a
method involving complex exponentials. It is possible to solve the equation without bringing
complex numbers into play, although the algebra is is perhaps slightly less obvious and depends
on applying some trigonometric identities.

Recall that the equation of motion is


γ F0
ẍ + ẋ + ω02 x = cos(ωe t) . (A1)
m m
We make the same assumption/guess as before about the form of the steady-state displacement:
xp (t) = A cos(ωe t − δ)
(A2)

which then implies


ẋp (t) = −ωe A sin(ωe t − δ)
(A3)
ẍp (t) = −ωe2 A cos(ωe t − δ)
for the velocity and acceleration.

Now, rather than re-expressing xp (t) as the real part of a complex-exponential function, we
can simply put equations (A2) and (A3) directly into equation (A1):

−ωe2 A cos(ωe t − δ) [ẍp ]


γ γ
− ωe A sin(ωe t − δ) [+ ẋ0 ]
m m
F0 F0
+ω02 A cos(ωe t − δ) = cos(ωe t) [+ω02 xp = cos(ωe t)]
m m

Collecting the terms in cos(ωe t − δ) on the left-hand side, this becomes


γ F0
ω02 − ωe2 A cos(ωe t − δ) − ωe A sin(ωe t − δ) =

cos(ωe t) . (A4)
m m
The main complication at this point is that the sinusoidal terms on the left-hand side of
equation (A4) take the argument (phase) (ωe t−δ), while the right-hand side only has the term
cos(ωe t) without the phase difference δ. To deal with this, we note that ωe t = (ωe t − δ) + δ,
so we can write
F0 F0
cos(ωe t) = cos [(ωe t − δ) + δ]
m m
F0
= [cos(ωe t − δ) cos δ − sin(ωe t − δ) sin δ]
m
F0 F0
= cos(ωe t − δ) cos δ − sin(ωe t − δ) sin δ .
m m
[The second line follows from the first because cos(Θ + Φ) = cos Θ cos Φ − sin Θ sin Φ for any
two angle Θ and φ.] Equation (A4) then reads
γ F0 F0
ω02 − ωe2 A cos(ωe t−δ)− ωe A sin(ωe t−δ) =

cos δ cos(ωe t−δ) − sin δ sin(ωe t−δ) .
m m m
– 22 –

Now, we need this equation to be true at any time t, and in order for this to happen the
term in cos(ωe t − δ) on the left-hand side must exactly equal the corresponding term on the
right-hand side; and separately the term in sin(ωe t − δ) on the left-hand side must exactly
equal the corresponding term on the right-hand side. That is, we require both
F0
ω02 − ωe2 A =

cos δ
m
γ F0
and ωe A = sin δ
m m
These are the same two equations involving the same two unknowns, A and δ, that we arrived
at from the complex-exponential method. The steps to solve for each of A and δ is the same
as the steps following equations (12) and (13) in Section 2.

B. Solutions for Forced Oscillators Including Transients

The following 6 figures show full solutions for the displacement of a forced, underdamped
oscillator, including the fast-decaying transient component as well as the steady-state solution
that we have been looking at in detail. For illustrative purposes, a value of γ = 0.125 mω0 <

2 mω0 has been chosen for all the calculations. Since this also satisfies γ < 2 mω0 , amplitude
resonance can occur in this case (see Section 4).

In each figure:

The top curve shows the time variation of F (t) = F0 cos(ωe t) for the harmonic external force,
for some assumed value (different from figure to figure) of the driving angular frequency ωe .

The second curve shows the steady-state solution to the equation of motion (6)—that is,
xp (t) = A cos(ωe t − δ), with the amplitude A and phase lag δ given in equations (14) and (15)
above.

The third curve shows the transient solution, which we have not derived explicitly here; but,
because we have set γ = 0.125 mω0 < 2 mω0 for these figures, the transient is of the basic form
discussed in Lectures 6–8 for an underdamped oscillator (i.e., oscillation with an exponentially
decaying amplitude).

The bottom curve in each figure shows the total motion x(t) of the driven oscillator, as the
direct sum of the steady-state and transient solutions. In all cases, the transient is important
only at relatively early times; eventually the steady-state xp (t) that we have derived dominates
completely.

The figures are not shown to scale, in the sense that no indication is given of the amplitude
A of the steady-state xp (t) in particular. This amplitude does change as a function of the
driving frequency ωe , for a given value of γ and the other constants (equation [14] and Figure
1). In particular, the displacement amplitude is much smaller at low and high ωe than it is at
resonance and near ωe = ω0 . However, the main emphasis here is on the overall shape of the
solution to the equation of motion, rather than on the full quantitative details.
– 23 –

Fig. 6.— Limit of low-frequency driving: with ωe = 0 exactly, F (t) = F0 cos(0) = F0 is


simply a constant external force. The steady-state xp (t) is therefore just a non-zero constant
[xp (t) = A = F0 /mω02 ], and so ultimately the driving force cancels the natural restoring
force and shifts the oscillator to a new equilibrium away from x = 0.

Fig. 7.— Still relatively low-frequency driving, but with a non-zero ωe = 0.1 ω0 , so the external
force does vary in time. The steady-state xp (t) is very nearly in phase with F (t) (the phase
lag δ is near zero), and thus so is the total motion once the transient has died away.
– 24 –

Fig. 8.— A somewhat larger driving angular frequency, ωe = 0.6 ω0 . Notice, in the total x(t),
the complicated interaction at early times between oscillations at the angular frequency of the
underdamped transient solution, and oscillations at the driving angular frequency. The latter
eventually dominate.

Fig. 9.— Velocity resonance: ωe = ω0 . The steady-state displacement xp (t) lags F (t) by
exactly 90◦ = π/2 rad = one-quarter cycle (i.e., δ = π/2), meaning that the steady-state
velocity is precisely in phase with the external force.
– 25 –

Fig. 10.— Motion for a driving frequency above the resonance frequency: ωe = 1.7 ω0 . Notice
that the steady-state xp (t) is clearly out of phase (but not quite in anti-phase) with F (t).
Notice also the complicated, two-frequency behaviour of the total x(t) at early times, before
the transient dies away completely.

Fig. 11.— Large ωe = 10 ω0 , towards the limit of high-frequency driving. The steady-state
displacement xp (t) is almost exactly in anti-phase with F (t) (the phase lag is δ ≃ π rad =
180◦ = one-half cycle). Here the driving force effectively replaces the natural restoring
force, giving simple harmonic oscillation at angular frequency ωe (with small amplitude).
– 26 –

C. Peaks of the Response Curves for x(t), ẋ(t), and ẍ(t) in the Steady State

Here we go through the algebra to derive the different driving angular frequencies that maximize
the amplitude of displacement, the amplitude of velocity, and the amplitude of acceleration
in the steady-state motion of a driven oscillator. These are the frequencies ωe at which the
response curves in Figure 3 peak.

C.1. Peak of the Displacement Response Curve: Amplitude Resonance

As shown in Section 2, the steady-state displacement amplitude is

F0 /m
A = p 2
(ω0 − ωe2 )2 + γ 2 ωe2 /m2

To find the driving frequency that maximizes A, we consider all of F0 , m, γ, and ω0 to be


constant, so that A is a function only of ωe . Thus, we need to find the derivative dA/dωe , set
it equal to 0, and solve for ωe . To do so, it is easiest to define a simpler function,

u(ωe ) ≡ (ω02 − ωe2 )2 + γ 2 ωe2 /m2

which contains the full dependence on ωe . Then we can write

A = (F0 /m) u−1/2

and differentiate this by the chain rule:


dA dA du
= ×
dωe du dωe
γ2
   
1 F0 −3/2 2 2
= − u × 2(ω0 − ωe ) × (−2ωe ) + 2 ωe 2
2m m
2
 
F0 −3/2 γ
= +ωe u × 2(ω02 − ωe2 ) −
m m2

Notice that in the second line the chain rule was also used to differentiate (ω02 − ωe2 )2 with
respect to ωe . The third line is a straightforward simplification of the second line. For the
derivative to be zero, it is clear that either ωe = 0 is required [corresponding to a constant
external force, F (t) = F0 cos(0) = F0 ] or the expression in the square brackets on the third line
must be zero. It is the second of these possibilities that leads to a maximum in A for a lightly
damped system (as could be confirmed by taking the second derivative, d2 A/dωe2 , although we
do not do this here). Thus,

γ2 γ2
maximum A =⇒ 2(ω02 − ωe2 ) = =⇒ ωe2 = ω02 −
m2 2 m2
This then gives the driving angular frequency,
q
ωe,max = ω02 − γ 2 /(2 m2 )

at which amplitude resonance occurs. Evidently, since γ ≥ 0 by definition, we have that


ω02 − γ 2 /(2m2 ) ≤ ω02 in general. Hence the steady-state displacement response curve
– 27 –

peaks at ωe,max ≤ ω0 in general. In addition, if ωe,max is to be non-zero and real, then


the quantity under the square root in its definition must be positive; that is, ω02 − γ 2 /(2m2 ) >
0. Therefore, amplitude resonance can occur only in systems that are sufficiently

underdamped, with γ < 2 mω0 . In systems with larger damping constants than this
(which may still be underdamped, so long as γ < 2 mω0 , just not sufficiently underdamped), the
displacement amplitude decreases monotonically as a function of increasing driving frequency
and there is no maximum of A away from ωe = 0.

C.2. Peak of the Velocity Response Curve: Velocity Resonance

The steady-state velocity amplitude is (see equation [16])

ωe F0 /m
ωe A = p 2
(ω0 − ωe2 )2 + γ 2 ωe2 /m2

Again, to find the driving frequency that maximizes the velocity amplitude (assuming that all
of F0 , m, γ, and ω0 are fixed), we need to find the derivative of (ωe A) and equate it to zero.
The product rule of differentiation tells us that
d dA
(ωe A) = A + ωe
dωe dωe

and then we can use our definition above of the function u(ωe ) ≡ (ω02 − ωe2 )2 + γ 2 ωe2 /m2 ,
together with the result above for dA/dωe , to write

γ2
 
d F0 −1/2 F0 −3/2 2 2
(ωe A) = u + ωe × ωe u 2(ω0 − ωe ) −
dωe m m m2
2
 
F0 −3/2 γωe
= u u + 2ωe2 (ω02 − ωe2 ) −
m m2
γ 2 ωe2 γωe2
 
F0 −3/2 2 2 2 2 2 2
= u (ω0 − ωe ) + + 2ωe (ω0 − ωe ) −
m m2 m2
F0 −3/2  2
(ω0 − ωe2 )2 + 2ωe2 (ω02 − ωe2 )

= u
m
Because the function u(ωe ) is never zero, the velocity amplitude is therefore maximized for the
driving angular frequency that makes the expression in square brackets vanish on the final line
here. That is,

maximum ωe A =⇒ (ω02 − ωe2 )2 = − 2ωe2 (ω02 − ωe2 )


=⇒ ω04 − 2ω02 ωe2 + ωe4 = − 2ωe2 ω02 + 2ωe4
=⇒ ω04 = ωe4

and hence
ωe = ω0
is the necessary condition for velocity resonance.
– 28 –

C.3. Peak of the Acceleration Response Curve

The steady-state acceleration amplitude is (see equation [16])


ωe2 F0 /m
ωe2 A = p 2
(ω0 − ωe2 )2 + γ 2 ωe2 /m2
Now we want to differentiate this quantity with respect to ωe and set the result equal to 0, to
identify the driving angular frequency that gives a maximum. By the product rule again,
d dA
(ωe2 A) = 2ωe A + ωe2
dωe dωe
Thus, bringing in the definition of u(ωe ) ≡ (ω02 − ωe2 )2 + γ 2 ωe2 /m2 and the result for dA/dωe ,
γ2
 
d 2 F0 −1/2 2 F0 −3/2 2 2
(ω A) = 2ωe u + ωe × ωe u 2(ω0 − ωe ) −
dωe e m m m2
γωe2
 
F0 −3/2
= ωe u 2u + 2ωe2 (ω02 − ωe2 ) −
m m2
2 2 γωe2
   
F0 −3/2 2 2 2 γ ωe 2 2 2
= ωe u 2 (ω0 − ωe ) + + 2ωe (ω0 − ωe ) −
m m2 m2
γ 2 ωe2 γ 2 ωe2
 
F0 −3/2 2 2 2 2 2 2
= ωe u 2(ω0 − ωe ) + 2ωe (ω0 − ωe ) + 2 2 −
m m m2
γ 2 ωe2
 
F0 −3/2
2(ω02 − ωe2 )( ω02 − ωe2 + ωe2 ) +

= ωe u
m m2
γ 2 ωe2
 
F0 −3/2
= ωe u 2ω02 (ω02 − ωe2 ) +
m m2
γ 2 ωe2
 
F0 −3/2 4 2 2
= ωe u 2ω0 − 2ω0 ωe +
m m2
γ2
  
F0 −3/2 4 2 2
= ωe u 2ω0 − 2ωe ω0 −
m 2m2
For this derivative to be zero requires, as before, that the expression in square brackets on the
last line be zero. Therefore,
γ2
 
2 4 2 2
maximum ωe A =⇒ 2ω0 = 2ωe ω0 −
2m2
ω04
=⇒ ωe2 = 2
ω0 − γ 2 /(2m2 )
and so we have .q
ωe = ω02 ω02 − γ 2 /(2m2 )
for the driving frequency that maximizes the steady-state acceleration amplitude.

This final expression can also be written as ω02 /ωe,max , where ωe,max is the driving frequency
that maximizes the displacement amplitude (amplitude resonance; Appendix C.1). Because
ωe,max ≤ ω0 in general, the maximization of acceleration amplitude occurs for driving

frequencies ≥ ω0 in general. Moreover, since γ < 2 mω0 is required for ωe,max to be
real, a peak in the steady-state acceleration response curve can occur only for sufficiently
underdamped systems that meet this condition (same as for the displacement response curve).
– 29 –

D. Averages of Trigonometric Functions Over an Oscillation Cycle

We are often interested in the average value of some property of an oscillator or a wave over
one full cycle of the oscillation (or over a whole number of cycles). For properties that vary
harmonically in time—that is, as some combination of sin(ω t + φ0 ) and cos(ω t + φ0 ), with a
well-defined angular frequency ω—one oscillation cycle corresponds to a change of exactly 2π
in the phase, (ω t + φ0 ). Thus, the mathematical results we need are the averages, between
θ = 0 and θ = 2π, of a few basic combinations of sin θ and cos θ. These are:
hsin θi = 0
hcos θi = 0
hsin2 θi = 1/2
hcos2 θi = 1/2
hsin θ cos θi = 0
where the angular brackets hi are used to denote the averaging over one cycle.

The following figure shows the curves of sin θ, cos θ, sin2 θ, and cos2 θ in the range 0 ≤ θ ≤ 2π
(ticks on the x-axes are at θ = π/2, θ = π, θ = 3π/2, and θ = 2π). It is easy to be convinced
from inspection of these that hsin θi = hcos θi = 0 and hsin2 θi = hcos2 θi = 1/2, as just stated.
We now prove these results rigorously, and also show directly that hsin θ cos θi = 0.

The proofs depend on understanding the definition of an average for a continuous function. In
general, the average of any function f (x) over any interval x1 ≤ x ≤ x2 is simply
R x2
f (x) dx
hf i = x1R x2
x1 dx
In our case, then, the averages of sin θ and cos θ are:
R 2π Z 2π
sin θ dθ 1 1 θ=2π 1
hsin θi = 0 R 2π = sin θ dθ = − [cos θ] θ=0 = − [1 − 1] = 0
dθ 2π 0 2π 2π
0
and similarly,
R 2π 2π
cos θ dθ 1 1 1
Z
θ=2π
0
hcos θi = R 2π = cos θ dθ = [sin θ] θ=0 = [0 − 0] = 0
dθ 2π 0 2π 2π
0
– 30 –

The averages of sin2 θ, cos2 θ, and the product (sin θ cos θ) follow from the standard relations

1 = cos2 θ + sin2 θ (D1)


2 2
cos(2θ) = cos θ − sin θ (D2)
sin(2θ) = 2 sin θ cos θ (D3)

To find hsin2 θi first, notice that subtracting equation (D2) from equation (D1) leads to
1
1 − cos(2θ) = 2 sin2 θ =⇒ sin2 θ = [1 − cos(2θ)]
2
Therefore,
R 2π
sin2 θ dθ 1 2π
1 12π
Z Z
2 0 2
hsin θi = R 2π = sin θ dθ = [1 − cos(2θ)] dθ
dθ 2π 0 2π 0 2
0
Z2π Z 2π 
1
= dθ − cos(2θ) dθ

 0 0

1 1 θ=2π
= 2π − [sin(2θ)]θ=0
4π 2
 
1 1 1
= 2π − [sin(4π) − sin(0)] = {2π − 0}
4π 2 4π
1
=
2
Similarly, to find hcos2 θi we add equation (D2) to equation (D1) to obtain
1
1 + cos(2θ) = 2 cos2 θ =⇒ cos2 θ = [1 + cos(2θ)]
2
and then,
R 2π
cos2 θ dθ 1 2π
1 1 2π
Z Z
2 0 2
hcos θi = R 2π = cos θ dθ = [1 + cos(2θ)] dθ
dθ 2π 0 2π
0 2
0
Z 2π 2π 
1 1
Z
= dθ + cos(2θ) dθ = {2π + 0}
4π 0 0 4π
1
=
2
1
Finally, given equation (D3), which says that sin θ cos θ = 2 sin(2θ), we have
R 2π 2π 2π
sin θ cos θ dθ 1 1 1
Z Z
0
hsin θ cos θi = R 2π = sin θ cos θ dθ = sin(2θ) dθ
2π 2π 2
0 dθ 0 0
 θ=2π
1 1 1
= − cos(2θ) = − [cos(4π) − cos(0)]
2π 4 θ=0

1
= − [1 − 1] = 0

Summary of Damped and Forced Oscillation (Lectures 6–10)

Equation of motion for a damped oscillator (undriven, or “free”):


γ
m ẍ + γ ẋ + mω02 x = 0 or ẍ + ẋ + ω02 x = 0
m
where
mẍ = net f orce ẍ = net acceleration
−γ ẋ = damping f orce −(γ/m) ẋ = damping accel0 n
−mω02 x = natural restoring f orce −ω02 x = restoring accel0 n

ω0 is the natural (or undamped) angular frequency. Associated natural period is


T0 = 2π/ω0 as usual.

x = 0 is equilibrium

Three classes of solution, depending on γ vs (2mω0 ) :

I. Underdamping/Light Damping γ < 2mω0 ⇐⇒ ω02 > γ 2 /(4m2 )


q
x(t) = A0 e−γ t/(2m) sin(ω t + φ0 ) with ω≡ ω02 − γ 2 /(4m2 )

• Motion is still oscillatory and periodic; return to equilibrium via overshoot and over-
correction

• ω as given is the damped angular frequency; period is T = 2π/ω ; in general:


2π 2π
ω < ω0 =⇒ T = = p 2 > T0
ω ω0 − γ 2 /(4m2 )

• A0 , φ0 : constants determined by initial conditions or other boundary conditions

• Damping force dissipates energy: mechanical −→ heat, etc; power is

Pdamp = Fdamp × ẋ = −γ ẋ2 = dEtot /dt

• Oscillation amplitude decays exponentially:


2m
A(t) = A0 e−γ t/(2m) =⇒ characteristic decay timescale :
γ

• Mechanical energy (kinetic plus potential) decays exponentially:


m
Etot (t) ∝ A(t)2 ∝ e−γ t/m =⇒ characteristic decay timescale :
γ
–2–

II. Critical Damping γ = 2mω0 ⇐⇒ γ/(2m) = ω0

x(t) = e−γ t/(2m) [C1 t + C2 ] = e−ω0 t [C1 t + C2 ]

• Motion is non-oscillatory and aperiodic

• C1 , C2 : constants determined by initial conditions or other boundary conditions

• Damping force dissipates energy, again at rate (power) Pdamp = −γ ẋ2 = dEtot /dt

• Monotonic approach to x = 0 at late times; fastest possible return to equilibrium

III. Overdamping/Heavy Damping γ > 2mω0 ⇐⇒ γ 2 /(4m2 ) > ω02

q
x(t) = e−γ t/(2m) B1 eq t + B2 e−q t γ 2 /(4m2 ) − ω02
£ ¤
with q≡

• Motion is non-oscillatory and aperiodic

• B1 , B2 : constants determined by initial conditions or other boundary conditions

• Damping force dissipates energy, again at rate (power) Pdamp = −γ ẋ2 = dEtot /dt

• Monotonic approach to x = 0 at late times; slower than for critical damping

• Larger γ =⇒ slower return to equilibrium (more resistance to any motion)


–3–

Equation of motion for a forced (or driven) oscillator harmonic F (t) ≡ F0 cos(ωe t) :
γ F0
m ẍ + γ ẋ + mω02 x = F0 cos(ωe t) or ẍ + ẋ + ω02 x = cos(ωe t)
m m

ωe is the driving (angular) frequency (subscript e = “external”)

Total motion x(t) is sum of two components:

transient : one of solutions above for undriven damped oscillators; independent of F 0 , ωe


dies away/decays relatively quickly

steady state : oscillation governed by external force (F0 , ωe ) as well as intrinsic γ, m, ω0


remains in the long term

Steady-state solution

x(t) = A cos(ωe t − δ) displacement amplitude = A


ẋ(t) = −ωe A sin(ωe t − δ) velocity amplitude = ωe A
ẍ(t) = −ωe2 A cos(ωe t − δ) acceleration amplitude = ωe2 A
Oscillation at the same frequency as the driving force, with a constant amplitude (A) ,
but possibly a phase lag (δ) in the response of x(t) to F (t)
F0 /m
A = p
2
(ω0 − ωe )2 + γ 2 ωe2 /m2
2

γωe /m
tan δ =
ω02 − ωe2

Low-frequency driving ωe ¿ ω0
δ → 0 as ωe → 0 ; steady-state displacement x(t) in phase with F (t) ; external force
works to cancel natural restoring force , shift oscillator to new equilibrium

High-frequency driving ωe À ω0
δ → π as ωe → ∞ ; steady-state displacement x(t) in anti-phase with F (t) ; external
force works to replace natural restoring force

ωe = ω0 (closely related to resonance)


δ = π/2 ; steady-state velocity ẋ(t) in phase with F (t) ; external force cancels damp-
ing force exactly ; “amplification” of sufficiently underdamped systems (γ < mω0 )

Resonance maximum interaction with external F (t); sharper for lighter damping (lower γ)
p
Amplitude Resonance : ωe,max = ω02 − γ 2 /(2m2 ) maximizes amplitude A = xmax

can only occur if γ < 2 mω0 (underdamped)
ωe,max < ω0 in general
Velocity Resonance : ωe = ω0 maximizes velocity amplitude, ωe A = ẋmax

Power input average hPext i = 12 γ ωe2 A2 = −hPdamp i at any ωe ; so net dEtot /dt = 0
ωe A is velocity amplitude; so power maximized at ωe = ω0

Response Curves see Lectures 9/10

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy