Waves Book
Waves Book
Fall 2024
Contents
6 Experiment 6: Diffraction 28
6.1 Purpose of the Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
6.2 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
6.3 Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
1
8 Experiment 8: Michelson Interferometer 39
8.1 Purpose of the Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
8.2 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
8.3 Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
2
1 | Experiment 1: Simple Harmonic Motion
To investigate simple harmonic motion using an oscillating spring and determine the spring constant of a
spring.
1.2 Introduction
As a model for simple harmonic motion, consider a block of mass m attached to the end of a spring Figure
1.1.
When the spring is neither stretched nor compressed, the block is at rest at the position called the equi-
librium position of the system, which we identify as x = 0. When the block is displaced to a position x,
the spring exerts on the block a force that is proportional to the position and given by Hooke’s law,
Fs = −kx (1.1)
where k is called the spring constant and is a measure of the stiffness of the spring. We call Fs a restoring force
because it is always directed toward the equilibrium position and therefore opposite the block’s displacement
from equilibrium. In such a case, simple harmonic motion occurs. Assuming no frictional forces and assuming
that the spring is massless, the equation of motion of a mass on a spring is,
d2 x
m = −kx (1.2)
dt2
or
d2 x k
=− x (1.3)
dt2 m
The solution of this second-order differential equation is,
s
k
x(t) = A cos(ωt + ϕ), ω= (1.4)
m
The period T, the frequency f, and ω are related by,
2π
ω = 2πf = (1.5)
T
3
Thus the period T is given by
r
m
T = 2π (1.6)
k
A very important property of simple harmonic motion is that the period T does not depend on the amplitude
of motion, A.
In this experiment, you will first determine the spring constant k of a spring, by hanging various weights
from the spring and measuring the extension. Then, you will measure the period of oscillation of a mass m
hanging from the spring and will compare this measured period with the period with the period predicted
by (1.6).
1.3 Experiment
• Hang the mass hanger from the spring and measure the initial height x0 .
x0 =
• Add 55 g to the mass hanger and determine the change in position caused by this added weight.
∆x1 =
k1 =
• Add 55 g masses incrementally until 275 g has been added to the mass hanger. Determine the total
displacement and the total added weight with each addition.
• Plot graph of F vs. ∆x using Graphical Analysis. Analyze the graph with a linear fit.
• Hang the mass hanger 55 g from the spring. Pull the mass hanger down slightly and release it to
create small oscillations. Measure the time required for 10 oscillations. (This is like measuring one
period ten times over.)
• Add 55 g masses incrementally until 275 g has been added to the mass hanger. Determine the period
with each addition.
4
Mass (g) Force (N) Duration (s) Period (s)
• Plot a graph of T 2 vs. m using Graphical Analysis. Analyze the graph with a linear fit. Calculate the
spring constant of the spring.
k1 = (1.7)
• Compare the result of Part 1 and Part 3. Calculate the error percentage by assuming the static case
gives the theoretical value of the spring constant.
5
6
2 | Experiment 2: Damped Harmonic Motion
2.2 Introduction
Newton’s Second Law for a system composed of two springs with constants k1 and k2 and a mass m is:
−
→
−
→ dP d dm d−
→v
F = = (m−
→
v)=−
→
v +m (2.1)
dt dt dt dt
As the mass is constant,
−
→ d−
→v
F =m = m−
→
a (2.2)
dt
The spring applies a restoring force that is always in the opposite direction to the displacement. The two
springs apply forces in the same direction so the effective spring constant is the sum of two,
−
→
F = k−
→
x , k = k1 + k2 (2.3)
The mass moves on a line in response to the restoring force. We can treat the motion as a one-dimensional
one. We take it as x.
dv d2 x
F = −kx = m =m 2 (2.4)
dt dt
This is a second-order differential equation with constant coefficients. Its solution is of the form; of x(t) = eαt ,
the two unknowns can be determined from the initial conditions. Substituting;
k
(mα2 + k)eαt = 0 −→ α2 = − (2.5)
m
This equation is known as the “characteristic equation” of the system. Since k and m are real, α must be
imaginary. The k
m ratio is the square of the “natural angular frequency” of the system, ω02 .
7
x(t) = Aeiω0 t + Be−iω0 t (2.8)
x = x0 cos(ω0 t + ϕ) (2.9)
Here x0 is the amplitude, ω0 is the angular frequency, and ϕ is the phase constant of the motion. (If at
t = 0, x = x0 then ϕ = 0.)
If there are frictional forces the amplitude of the motion gradually decreases. Such a motion is known as
DHM. When the damping force is proportional to the speed of the block, equation 2.1 is replaced by,
!
d2 x dx
m +b + kx = 0 (2.10)
dt2 dt
where b is the damping coefficient.
Question 2: Solve (2.10) assuming x(t) = eαt : Obtain the characteristic equation for α, find α itself and
obtain the function x(t) given below:
2 !1/2
2k b
ω= − (2.11)
m 2m
b
x(t) = A0 e− 2m t cos(ωt + ϕ) (2.12)
Note that the constant 2k represents the total spring constant of the two-spring system. Here, the amplitude
b
of the harmonic function is decreasing exponentially as A = A0 e− 2m t the time that elapses for A to decrease
to 1/e of its original value is A0 is known as the relaxation time(durulma zamani).
τ = 2m/b (2.13)
2
If b
2m < 2k
m is satisfied the system is said to be underdamped. The angular frequency, ω, is less than the
underdamped case. The motion is shown in Figure 2.2
8
2 2
If b
2m > 2k
m the system is overdamped, and if b
2m = 2k
m the system critically damped. In overdamped
and critically damped cases there are no oscillations.
Question 3: Explain mathematically why we would not observe oscillations in the overdamped and critically
damped cases. Which form does x(t) assume in these cases; i.e., what form does (2.12) take?
Question 4: Research in which cases the use of a linear (in speed) term for the frictional force is justified
and in which cases should one switch to a term proportional to the speed squared, i.e. v 2 . Think of general
criteria for the parameters of the system, such as its size.
2.3 Experiment
• Place first the carbon paper and then the white paper on the air table. Place the gliders as seen in
the picture.
• Start the air pump which causes the gliders to slide easily on the table. The spark generator should
be set at an appropriate frequency. Try to pull the glider in the direction of the springs so that the
motion will remain one-dimensional.
• Press the pedal of the spark generator and simultaneously release the glider. Note that from this
moment on there is high voltage on the chains, on the glider hoses, on the metal parts of the gliders
and the paper on the table.
• Pull the blank paper you placed on the carbon paper slowly with constant speed until the glider comes
to rest. You will see the trace of the motion of the glider on the reverse side of this paper.
Data Analysis
• Mark the points of maximum displacement for each oscillation; find their values and logarithms of
these values. Fill the table with your data. (The sparks are generated at equal intervals. Thus the
number of dots gives the time between two events.)
9
t (.....) A (.....) lnA(.....)
• Plot (lnA versus t). Is this a straight line? If it is a line write its equation as;
• Find the relaxation time and the damping coefficient. Do your results agree with the ones obtained
in part A of the experiment?
Questions
• What are the possible causes of error in this experiment and how can they be reduced? Calculate the
errors in τ and b.
10
11
3 | Experiment 3: Coupled Oscillations
3.2 Introduction
Many important physics systems involve coupled oscillators. Coupled oscillators are oscillators connected in
such a way that energy can be transferred between them. The motion of coupled oscillators can be complex
and does not have to be periodic. However, when the oscillators carry out complex motion, we can find a
coordinate frame in which each oscillator oscillates with a well-defined frequency.
A solid is a good example of a system that can be described in terms of coupled oscillations. The atoms
oscillate around their equilibrium positions and the interaction between the atoms is responsible for the
coupling. To start the study of coupled oscillations we will assume that the forces involved are spring-like
forces i.e. the magnitude of the force is proportional to the magnitude of the displacement from equilibrium.
Two simple pendulums can be connected with a spring to obtain coupled pendulums as shown in Figure 3.1.
Pendulums are identical, having the same mass M and length l. The deflection angles are small so that the
small angle approximation can be used i.e. sin θ ∼
= θ. The force acting on the pendulum a in the direction
of motion due to the coupling is −Kl (θa − θb ) and the force acting on the pendulum b is Kl (θa − θb ). Also,
their equations of motion are:
d2 θa g K
2
+ θa + (θa − θb ) = 0 And
dt l M (3.1)
d2 θb g K
2
+ θb − (θa − θb ) = 0
dt l M
These equations can be put into the matrix form:
g
+ K K
−M θa θ̈a
l M = − (3.2)
g
K
−M l + K
M θb θ̈b
12
In order to find non-trivial solutions, the determinant of the operating matrix should be zero i.e.
2
g K K2
+ − ω2 − =0 (3.5)
l M M2
g g K
ω12 = & ω22 = +2 (3.6)
l l M
Then, it is easy to find relations between θa and θb . For example, let us choose ω = ω1 , from (3.4) we see
that
K K
−M θa
M =0 (3.7)
K K
−M M θb
Figure 3.2. Initial positions of pendulums for the symmetric mode (left), the anti-symmetric mode (middle)
and the beat (right).
These two cases correspond to normal modes and ω1 , ω2 are called normal frequencies. In a normal mode,
both objects move with the same frequency and the phase constant. The former case is the symmetric mode
and the latter case is the anti-symmetric mode as it has been illustrated in Figure 3.2. The symmetric mode
has lower energy than the anti-symmetric mode since higher frequencies correspond to higher energies.
Therefore, the general solution of the system can be written as:
θa (t) =A1 cos (ω1 t) + B1 cos (ω2 t) + A2 sin (ω1 t) + B2 sin (ω2 t)
(3.8)
θb (t) =A1 cos (ω1 t) − B1 cos (ω2 t) + A2 sin (ω1 t) − B2 sin (ω2 t)
Then the arbitrary coefficients in (3.8) turn out to be as A2 = B2 = 0 while A1 = B1 = A/2. Using
trigonometric identities (3.8) becomes
ω1 + ω2 ω1 − ω2
θa (t) =A cos t cos t
2 2
(3.10)
ω1 + ω2 ω1 − ω2
θb (t) =A sin t sin t
2 2
In that case, the objects move as shown in Figure 3.3. This phenomenon is called beat. Energy is exchanged
continuously between pendulums while the system moves with the beat frequency.
13
Figure 3.3. Beat: ∆ω = (ω1 − ω2 )/2.
3.3 Experiment
• Using Hooke’s Law, find the spring constant of the coupling spring.
κ=
• Measure the time of 5 periods in the symmetric mode and find the period.
• Repeat the experiment 5 times and fill the table below for symmetric mode.
• Measure the time of 5 periods in the asymmetric mode and find the period.
• Repeat the experiment 5 times and fill the table below for anti-symmetric mode
14
• Find the experimental frequencies from the average periods.
ω1exp = ω2exp =
ω1theo = ω2theo =
• Compare the experimental and theoretical frequency values by making an error analysis.
PART 2. Beat
ω exp =
ω theo =
• Compare the experimental and theoretical frequency values by making an error analysis.
15
16
4 | Experiment 4: Waves on a String
To observe standing waves on a stretched string and to determine the frequency of the line current.
4.2 Introduction
Let us first review the basic mathematics for the description of waves. The general form for the function
representing a travelling wave is y = f (x ± vt). The following expressions all depict different travelling
waves:
y = A sin(x − vt) (4.1)
y = ex−vt (4.3)
(Note that subtracting/adding a number from/to the argument of a function causes its curve to be shifted
to the right/left)
Question 1: Investigate “wave equation” and derive it by applying Newton’s second law to an infinites-
imal piece of wire under a tension T. Show by explicit substitution that all the functions given above satisfy
it.
Question 2: Show that
y = A sin/cos[k(x ± vt)] (4.4)
is a traveling harmonic wave (here sin/cos means “sin or cos”). What is the direction the wave is travelling
depending on the +/− in the expression?
These waves are periodic. The wavelength λ of a periodic wave is the distance between two points of equal
phase. That is in the function shifting x by λ should not change anything in the graph.
Note that changing the argument of a sin or cos function by 2π gives the same function value.
λ
v= (4.7)
T
Transverse sine waves are shown in Figure 4.1. They can be thought of as the superposition of an incoming
wave and the wave reflected from the impenetrable barrier at x = 0. The fact that the barrier is impenetrable
implies that the wave gets reflected without any distortion in its shape and without any energy loss.
Question 3: Find the superposition of the two waves (the incoming one and the reflected one) using the
trigonometric identity:
1 1
sin α + sin β = 2 sin[ (α + β)] cos[ (α − β)] (4.8)
2 2
The wave thus formed is known as the “standing wave”.
17
Figure 4.1. Time dependence of standing waves.
Here, 2A sin(kx) is the amplitude of this wave. Note that the amplitude of oscillation of any point depends
on its position. At points where |sin(kx)|= 1, that is kx = 2 ,
nπ
where n is an odd integer, the amplitude of
motion is maximum. Such points are known as antinodes (‘karın noktaları’). Their positions are given
by (using k = λ ):
2π
2n + 1 1 λ
xn = λ= n+ (4.10)
4 2 2
So, the distance between adjacent antinodes, which are denoted in the figure with K, is half a wavelength.
At kx = 0, π, 2π, ...|sin(kx)|= 0 and the string is stationary. These points (denoted by D in the figure) are
the nodes (‘düğüm noktaları’), and their positions are
λ
xm = m (4.11)
2
The distance between adjacent nodes is also half a wavelength. The two edges of the string are naturally
nodes if the string is clamped at its two edges in the experiment and as a consequence, an integer number
of half wavelengths fit between the boundaries of the string of length L.
λ
L=n (4.12)
2
Thus a standing wave can have only definite possible wavelengths which satisfy the boundary conditions.
The relation among its speed, frequency and wavelength is:
v = λf (4.13)
Substituting we obtain
v v
fn = =n (4.14)
λn 2L
These are the natural frequencies of the system. The lowest natural frequency, f1 = v
2L is known as the
fundemental (‘temel frekans’). For a wire of linear mass density µ = M
L and under a tension F = T ,
18
q q
the wave speed is v = T
µ and hence the fundamental frequency is f1 = 1
2L µ.
T
The other frequencies of the system are known as the harmonics and are given by:
s
1 T
fn = n (4.15)
2L µ
T = λ2 f 2 µ (4.16)
4.3 Experiment
The experimental setup in Figure 4.2 includes a signal producer and a wave producer. The latter converts
the electrical signal produced by the former into mechanical vibrations and transmits them to the string.
The tension in the string and the wave speed are adjusted by changing the hanging masses.
When the frequency driving the string is close to the natural frequency of the string we have a resonance
and observe standing waves.
• Start your signal and wave producers and increase the tension until you find a single half wave in your
string.
• In resonance, the amplitude is large and it is easily observable. Measure the wavelengths for L = λ/2
and L = λ using a ruler.
• Repeat the process above with different masses and fill the table on the next page.
19
• Plot T = f (λ2 ) and find its slope:
Slope = ...... (4.17)
• Find the frequency of the signal by equating the slope to f 2 µ. The linear mass density of the wire and
the frequency of the wave:
µ = ..... (4.18)
f = ..... (4.19)
Questions
• Find the numerical values of wave speed for the least and greatest tensions in the string and wire.
• What changes did using a string instead of a wire cause in this experiment?
• How does the wave speed change if the tension is (i) increased to 8 times and (ii) reduced to 1/3 of
its original value?
20
21
5 | Experiment 5: Resonance in an RLC Circuit
To observe and analyze the resonance phenomenon in a driven series RLC circuit, and understand the effect
of resistance on resonance sharpness.
5.2 Introduction
In an ideal LC circuit, resonance occurs when the inductive reactance (XL = ωL) equals the capacitive reac-
tance (XC = 1/ωC), allowing maximum energy transfer between the inductor and capacitor. In a practical
scenario, resistance, such as the internal resistance of a signal generator, creates an RLC circuit, which
broadens the resonance peak. The goal is to observe this resonance using an oscilloscope by varying the
frequency of the input AC signal.
1
ω0 = √
LC
At this frequency, the energy alternates between the magnetic field of the inductor and the electric field
of the capacitor, creating the resonance effect. However, in practical circuits, the presence of resistance
(R) causes energy dissipation, resulting
in a broadened and less sharp resonance peak. This resistance also
q
affects the circuit’s quality factor Q = R C , with a higher resistance leading to a lower Q-factor and a
1 L
In terms of the voltage gains (G), which are the output amplitudes divided by the input amplitude, the
resonance frequencies for the inductor, ωL , and capacitor, ωC , in the high-Q case can be approximated by:
1
ωL,C ≈ ω0 1±
4Q2
At these resonance frequencies, the voltage gains across the inductor and capacitor are approximately equal
22
to the quality factor Q1 :
The voltage gain as a function of the input frequency varies as shown in the transmissibility curves below,
with a peak at the resonance frequencies.
Figure 5.2. The transmissibility curves showing voltage gains for Q = 3 against the driving frequency
(dimensionless) ω. The resonance frequencies ωL and ωC are close to but not exactly equal to ω0 . The
resonance gains GL and GC are close to but not exactly equal to Q. Note the unit gains on L at high input
frequency and on C at low input frequency.
In a driven RLC circuit, by measuring the resonant frequencies ωL and ωC on an oscilloscope we can
determine ω0 and Q as:
1
ω0 = (ωL · ωC ) 2
− 1
ωC
2
Q= 2 1−
ωL
From the calculated Q, knowing the values of the inductance L and the capacitance C, the circuit’s resistance
R can then be determined using the following formula:
s
1 L
R=
Q C
5.3 Experiment
The experimental setup consists of a series LC circuit, driven by a signal generator, with an oscilloscope
used to observe the output signals. In our experiment, the inductance is L = 33 mH and the capacitance
is C = 22 nF. Although no explicit resistor is included, the circuit has internal resistance, primarily due to
the internal resistance of the signal generator.
1
The exact relations are:
± 12
1
ωL,C = ω0 1−
2Q2
− 12
1
GL,C =Q 1−
4Q2
23
PART 1. Resonance in RLC oscillator
• Connect the inductor and capacitor in series with the signal generator. Set the signal generator to
output a sinusoidal AC signal. Keep the amplitude knob on the signal generator fixed throughout the
experiment. The internal circuitry of the signal generator includes transistors or operational amplifiers
that behave differently at different signal levels, influencing the output impedance. Higher output levels
may push these components closer to their limits, resulting in changes in their effective resistance.
• Calibrate the oscilloscope first. Then, connect the oscilloscope probes across the inductor. Adjust
the oscilloscope settings to properly display the sinusoidal waveform. Vary the frequency of the signal
generator to locate and note the resonance frequency fL where the voltage across L peaks.
• Observe the sinusoidal output on the oscilloscope. Use the oscilloscope’s time scale to measure the
frequency of the waveform.
• Move the oscilloscope probes across the capacitor. Observe the sinusoidal output and measure the
frequency as before. Vary the frequency of the signal generator to locate and note the resonance
frequency fC where the voltage across C peaks.
h i−1/2
• Using these readings of fL and fC calculate f0 = (fL · fC )1/2 , and Q = 2 1 − fC
fL .
• From these results calculate the best value of f0qwith its uncertainty. Compare this with the resonant
frequency f0 derived from L and C by f0 = 2π 1
LC .
1
• Evaluate the discrepancy between the two values for f0 . Consider the uncertainties of both f0 ’s, as
the significance of the discrepancy depends on the uncertainties.
• Calculate the best value of Q with its uncertainty. Since the gains in the inductor and the capacitor
at resonance have a value close to Q, measure the gains at resonance and compare these with the
calculated value for Q. One way to measure the gains at resonance is to measure the input and output
amplitudes directly via the oscilloscope probe. Alternatively, compare the resonance gain of L with
its high-frequency gain, and the resonance gain of C with its low-frequency gain. The transmissibility
plot above testifies for the viability of this method.
• From the best value of Q, calculate the resistance R of the circuit. Compare this value of resistance
with a direct measurement via a digital multimeter (DMM) across the signal generator. Note that
the readings on the DMM may oscillate even in the correct mode. The reason is that the presence of
AC signals can cause the multimeter to encounter a dynamic load, and since DMMs typically measure
resistance using a small DC current, the AC signal can introduce voltage variations that affect the
readings.
24
PART 2. Lissajous curves
Lissajous curves are complex, closed-loop figures that are formed when two sinusoidal oscillations are com-
bined. They are typically displayed on an oscilloscope, where the X and Y axes represent two different
signals, often at right angles to each other. Lissajous curves offer a way to visualize the interaction between
two oscillating signals, revealing their amplitude, frequency, and phase relationships.
Lissajous curves can be described mathematically using the parametric equations x(t) = A sin at + δ, and
y(t) = B sin bt, where:
• a and b are the frequencies of the oscillations in the X and Y directions, respectively.
Lissajous curves are used in oscilloscopes to analyze the relationship between two periodic signals, especially
in the calibration of oscilloscopes. They also help visualize phase differences and amplitude ratios between
signals, useful in electronics and physics.
• Connect two function generators to the X and Y inputs of the oscilloscope using BNC cables.
• Set the oscilloscope to the dual-channel mode (XY-mode) to display both inputs simultaneously.
Adjust the time base and vertical scales to ensure the waveforms are properly displayed.
• Use the oscilloscope controls to adjust the trigger level and mode until you can clearly see the Lissajous
figure on the screen.
• Change the frequency ratio and phase difference of the signals to explore various Lissajous shapes.
Record observations of the curves produced for different settings.
25
• Now, probe the voltages across L and across C for the two channels of the oscilloscope. At resonance,
observe the shape of the Lissajous curve. Considering the amplitudes, frequencies, and the phase
difference of the input signals derived from the equations, comment on the shape of the Lissajous
curve. Is the shape on the oscilloscope what you expect?
PART 3. Questions
• How do the resistance, inductance, and capacitance components degrade over time in practice? In
particular, how would their values change; do they increase or decrease? Why?
• How would you measure the capacitance of a capacitor using an RC circuit? How would you measure
the inductance of an inductor using an RL circuit? Explain.
• How does the quality factor Q affect the sharpness of the resonance peak in a driven series RLC
circuit? What role does the internal resistance of the signal generator play in the overall behavior of
the RLC circuit?
• In our experiment, the internal resistance of the signal generator was significant. How could you adjust
the values of L and C to make the internal resistance of the function generator negligible?
• In what ways do Lissajous curves provide insight into the relationship between the voltages across the
inductor and capacitor in a driven RLC circuit?
• How would the behavior of the circuit change if we switched from a series RLC configuration to a
parallel RLC circuit, and how would this affect the resonance and impedance characteristics?
26
27
6 | Experiment 6: Diffraction
To observe and study the diffraction of light from single, double, and multiple slits. The wavelength of the
monochromatic light, as well as the slit separation distance of the double- and multiple-slit gratings, will be
calculated.
6.2 Introduction
Figure 6.1. Top: Superposition of two waves that are “in phase”. Bottom: Superposition of two waves
that are completely “out of phase”.
Figure 6.1 shows two sets of sinusoidal waves and their superpositions at time t = 0. In the top panel of Figure
6.1, the two waves “in phase”, Φ1 = sin (x) and Φ2 = 2 sin (x), form constructive interference whereas in the
bottom panel of Figure 6.1, the waves that are completely “out of phase”, Φ1 = sin (x) and Φ2 = 2 sin (x + π),
form destructive interference. Given the two sinusoidal functions Φ1 = A1 sin (kx − ωt + ϕ1 ) and Φ2 =
A2 sin (kx − ωt + ϕ2 ), the condition ϕ1 − ϕ2 = 0, ±2π, ±4π, · · · , ±2nπ, (n ∈ I) ensures that the waves are in
phase and ϕ1 − ϕ2 = ±π, ±3π, · · · , ±(2n − 1)π, (n ∈ I) means they are completely out of phase.
Run the P ython code given below and investigate the superposition of waves created by two-point sources.
Note that the graphics show the wave intensity (squared amplitude) rather than the amplitude. You will
observe that at some points the function oscillates while at others, where destructive interference occurs,
it remains constant. The system simulated by the code is analogous to plane waves passing through two
pinholes or slits shown in Figure 6.4.
When monochromatic light passes through a double-slit, in which the openings act as individual sources of
coherent light, interfering waves form light and dark fringes on the screen, forming an interference pattern.
Bright fringes correspond to interference maxima and show that waves arrive at the screen in phase. On
the contrary, dark fringes are interference minima and show that waves arrive at the screen out of phase.
28
Figure 6.2
Figure 6.3. A snapshot from the movie showing the interference pattern due to the two-point sources.
If the experiment was performed with tomatoes instead of light waves, we would observe two bunches of
29
splattered tomatoes directly ahead of the slits. Even if a spread were to occur as tomatoes pick up transverse
momentum bouncing off the edges of the slits, destructive interference would never take place for particles
with no wave properties.
6.3 Experiment
Considering an ideal double-slit setup in which the openings are infinitely small, the incident waves produce
the interference pattern given by Figure 6.5 on the screen. For coherent sources, as is the case in the given
configuration, the total intensity resulting from constructive interference is equal to four times the intensity
of a single source, i.e. Itot = 4I1 . For destructive interference, Itot = 0. In Figure (6.4), the wave passing
through the lower slit travels an extra distance δ towards point P. Hence, at this point, the electric field
components of the plane waves coming from the upper and lower slits may be set to
ϕ ϕ ϕ
Itot ∝ ⟨E ⟩ =
2
4E02 cos2 ⟨sin2 kx − ωt + ⟩ = 2E02 cos2 (6.2)
2 2 2
In terms of the maximum intensity I0 , which denotes the intensity at the locations of bright fringes in Figure
6.5, the above expression may be expressed as
ϕ
Itot = I0 cos 2
(6.3)
2
Provided that the distance L from the slits to the screen in Figure 6.4 is much greater than the slit width d,
the distance δ may be approximated as δ ≈ d sin(θ). Using this together with the relation δ = ϕ/k deduced
from constructive interference condition, (6.3) becomes
kd sin(θ)
Itot = I0 cos2 (6.4)
2
30
• In our setup, we are using a Helium-Neon (He-Ne) laser which emits red light with a wavelength of
632.8 nm.
• Place the double-slit in front of the He-Ne laser and measure the distance L from the double-slit
grating to the screen.
• Measure also the distance between the symmetric n-th dark fringes of your choice (2y) on the screen.
Keep in mind that the double-slit experiment in the previous section is performed width slits of finite
width, i.e. they are not infinitesimal openings, thus the intensity pattern on the screen is confined to
the envelope seen in Figure 6.5, which is a consequence of single-slit diffraction.
• Using the condition δ = (n + 12 )λ, (n ∈ I) for destructive interference, calculate the slit separation
distance d of the double-slit grating. (Hint: δ ≈ d sin(θ) ≈ dy/L.)
In order to formulate the interference pattern in Figure 6.6, it is useful to split the opening into two parts of
equal width. At the location of the first dark fringe on the screen, the waves coming from sources that are
b/2 apart across the slit should be completely out of phase. Namely, if there are n point sources in each half,
the contribution of the 1st source should cancel that of the (n + 1)th source and this pairwise cancellation
should apply to all the points up to the nth and 2nth sources. Consequently, for the first dark fringe, one
finds δ = b
2 sin(θ) and this yields the condition
b λ
sin(θ) = (6.5)
2 2
for destructive interference. For nth minimum, i.e. the nth dark fringe, the relation takes on the form
31
As for the intensity distribution, dividing the slit into N segments of width ∆x << λ, and considering
the rays traveling towards the same point P on the screen to be parallel to one another (as in Figure 6.6),
one finds that waves emerging from any two adjacent segments have a phase difference of ϕ = k∆x sin(θ).
Accordingly, the electric field components to be summed have the forms
E1 =C sin(ωt)
E2 =C sin(ωt + k∆x sin(θ))
.. (6.7)
.
EN =C sin(ωt + (N − 1)k∆x sin(θ))
After some algebra, the final expression for total electric field E = E1 + E2 + · · · + EN may be cast into the
form
sin(kb sin(θ)/2)
E=C sin(ωt + (N − 1)k∆x sin(θ)/2) (6.8)
sin(k∆x sin(θ)/2)
The total intensity follows as
2
C 2 sin(kb sin(θ)/2)
Itot ∝ ⟨E ⟩ =
2
(6.9)
2 sin(k∆x sin(θ)/2)
and as ϕ → 0, in terms of the maximum intensity I0 , it becomes
2
sin(kb sin(θ)/2)
Itot = I0 (6.10)
k∆x sin(θ)/2
Figure 6.6 shows the light intensity pattern for single-slit diffraction. As mentioned earlier, dark fringes obey
(6.6). In the small-angle approximation, the width of the fringes is given by sin(θ) ≈ tan(θ) = y/L = nλ/b
and thus the width of the central fringe is ∆ycenter = 2λL/b. Note that the fringe width is inversely
proportional to the slit width.
• Place the single-slit in front of the He-Ne laser and measure the distance L from the double-slit plate
to the screen.
• Measure the width of the central bright fringe (the distance between the symmetric first dark fringes).
• Using the formula for the central fringe width (∆ycenter = 2λL/b), verify the wavelength of the He-Ne
laser, that is λ = 632.8 nm.
Locations of the peaks are formulated via d sin(θ) = nλ and they get narrower with an increasing number of
slits. As demonstrated in Figure 6.7, when the beam consists of multiple wavelengths, narrow peaks allow
easy distinguishing of their separate diffraction patterns.
• Once you observe the interference pattern, repeat the previous measurements to find the slit-separation
distance d.
32
Figure 6.7. Intensity distribution of double-slit interference for infinitesimal slits (cyan line) versus slits of
finite width (blue line). The envelope (brown line) represents the single-slit diffraction effect coming from
the individual openings.
Questions
• How is the double-slit intensity formula in (6.4) modified for slits of finite-width b?
• What would you expect to see in the single-slit experiment for infinitesimally thin slits?
33
34
7 | Experiment 7: Polarization of Electromagnetic Waves
To observe the polarization of electromagnetic waves (EMW) and measure some polarization properties.
7.2 Introduction
In EMW the electric and magnetic fields are perpendicular to each other and to the direction of propagation
of the wave. The simplest polarization state is linear polarization. For the E field in the z and the B field in
the y directions, the waves propagate in the E × B or y direction. Traditionally the direction of E is taken
as the polarization of the wave. The most general form of the E field (and thus the polarization) is given by
The E field of the wave described here has both x and z components. Here ϕx and ϕz depend on our reference
axes but ϕx − ϕz is independent of our choice. Linear polarization corresponds to ϕx = ϕz . Another special
case is circular polarization: ϕx − ϕz = π/2 and Ex = Ez . When microwaves impinge on a metal grill
the component of the electric field parallel to the grid wires or bars is either absorbed or reflected because
electrons can move freely in this direction. The component of the electric field perpendicular to the grid
is not affected. It is as if the grid is an insulator in this direction. Thus only microwaves with an E field
perpendicular to the grid pass and the wave is polarized. (In the case of visible light polymeric material with
molecules aligned along a direction serves the same purpose as the grid.) The grid is known as a polarizer.
Figure 7.1
If a polarized EM wave is passed through another grid the E is only the field component perpendicular to
the second grid passes through. The light is now polarized by the second grid. Because the second grid is
used to analyze the polarization state it is called the analyzer. When light passes through a polarizer and
an analyzer the intensity of the light passing through the apparatus is proportional to the square of the E
field which is proportional to the cos of the angle between the polarizer and the analyzer. Thus the Malus
law
I = I0 cos2 (θ) (7.2)
35
Figure 7.2
7.3 Experiment
PART 1.
• Place the microwave (MW) transmitter and the receiver opposite to each other as in Figure 7.2. These
devices produce polarized MWs and are sensitive to MWs of one polarization state. Demonstrate this
by rotating one device 900 . What does the ammeter show?
• Set the receiver and transmitter as in Figure 7.2. Rotate the receiver by 100 intervals and fill in the
table.
• Plot I/I0 = f (θ) vs. θ and compare the graphic with cos2 (θ).
PART 2.
Figure 7.3
• Place a polarization grid between the polarizer and the analyzer in Figure 7.2. What do you observe
when the bars of the grill are parallel and perpendicular to the polarization plane? Why?
• Now rotate the receiver by 900 . What do you observe? Now place the grid as in Figure 7.3. Rotate it
as before. What do you observe? Why?
36
PART 3.
Figure 7.4
• Set up the equipment as in Figure 7.4. Place the transmitter at 450 angle. Place the grill and the
metal plate (reflector) as in Figure 7.4. You can obtain circularly polarized waves by adjusting the
distance between the grill and the plate.
• To check if you have circular polarization rotate the receiver through a full circle. When the value you
read on the receiver is independent of the angle the radiation is circularly polarized. (How did I come
to this conclusion?) Measure the distance between the grill and the plate and find at which distances
the wave is linearly and at which distances are circularly polarized.
37
38
8 | Experiment 8: Michelson Interferometer
To investigate how a Michelson interferometer works and determine the wavelength used.
8.2 Introduction
Instruments using the interference of waves to perform measurements are known as interferometers. Since
visible light has a very short wavelength distances can be very accurately measured using visible interferome-
ters. Michel- son interferometer, probably the most famous interferometer was developed in 1880. Michelson
has used his interferometer to determine the wavelength of cadmium to the accuracy of one in 108 . He later
showed that the speed of light is independent of the direction of travel which led to the theory of relativity.
Light from the source (S) is partially reflected and partially passes through the semi-reflecting mirror (BS).
(The thickness of the silver coating on this mirror is adjusted to reflect half of the light.) At his moment
both the reflected and passing light waves are in the same phase. The reflected beam is reflected from the
movable mirror and the reflected light is from the fixed mirror. Both are reflected to the semi-reflecting
mirror and continue to the screen. The two beams have travelled different distances. The superposition gives
a maximum (constructive interference) if the path difference is an integer multiple of full wavelength and
minimum (destructive interference if it is a half-integer multiple of a full wavelength. (Note the conceptual
similarity to double-slit interference. Note also that the distance between the mirrors is traversed twice so
that a mirror must be moved a half (not full) wavelength to go from one maximum to the next or one
minimum to the next.)
As the mirror is moved the interference pattern shifts, when it is moved a distance d the path difference
changes by d. If this equals the wavelength the pattern reverts to the original,
Wavelength is determined by noting how many times the pattern repeats itself when the mirror is moved a
distance d.
8.3 Experiment
(Be extremely careful when working with the interferometer. Even the slightest movement of the set destroys
or changes the patterns. You can see this by pressing your finger on the experiment table.)
• Place the electromagnetic wave receiver, transmitter, mirrors, and the grid as shown in Figure 8.1.
Note that, mirrors should be placed at the same distance from the grid.
• Move the mirror until the ammeter measures almost zero intensity (destructive interference).
• Continue to move the mirror until observing five consequent minimums (deconstructive interference)
and record the displacement of the mirror between each minimum in the table below.
39
Figure 8.1. Schematic representation of the setup of the experiment.
d λ
• Calculate wavelengths by using the relation given in (8.1). The experimental wavelength is the mean
of these five wavelengths.
λexp = ... (8.2)
Questions
40
41