You are on page 1of 30

1

Chapter 1
Interferometry and Material Introduction

1.1 Introduction to interference

In physics, interference is a phenomenon in which two waves superimpose to


form a resultant wave of greater or lower amplitude. Interference usually refers to the interaction
of waves that are correlated or coherent with each other, either because they come from the same
source or because they have the same or nearly the same frequency. Interference effects can be
observed with all types of waves, for example, light, radio, acoustic, matter waves and surface
water waves.

1.2 Mechanism of interference

The principle of superposition of waves states that when two or more waves are incident on the
same point, the total displacement at that point is equal to the vector sum of the displacements of
the individual waves. If a crest of a wave meets a crest of another wave of the same frequency at
the same point, then the magnitude of the displacement is the sum of the individual magnitudes –
this is constructive interference. If a crest of one wave meets a trough of another wave then the
magnitude of the displacements is equal to the difference in the individual magnitudes – this is
known as destructive interference.

Resultant wave

Wave 1

Wave 2

Fig.1.1Constructive and Destructive interference


2

Constructive interference occurs when the phase difference between the waves is a multiple of 2π,
whereas destructive interference occurs when the difference is π, 3π, 5π, etc. If the difference
between the phases is intermediate between these two extremes, then the magnitude of the
displacement of the summed waves lies between the minimum and maximum values.

Consider, for example, what happens when two identical stones are dropped into a still pool of
water at different locations. Each stone generates a circular wave propagating outwards from the
point where the stone was dropped. When the two waves overlap, the net displacement at a
particular point is the sum of the displacements of the individual waves. At some points, these will
be in phase, and will produce a maximum displacement. In other places, the waves will be in anti-
phase, and there will be no net displacement at these points. Thus, parts of the surface will be
stationary—these are seen in the figure above and to the right as stationary blue-green lines
radiating from the centre.

 Between two plane Waves

A simple form of interference pattern is obtained if two plane waves of the same frequency
intersect at an angle. Interference is essentially an energy redistribution process. The energy which
is lost at the destructive interference is regained at the constructive interference. One wave is
travelling horizontally, and the other is travelling downwards at an angle θ to the first wave.
Assuming that the two waves are in phase at the point B, then the relative phase changes along
the x-axis. The phase difference at the point A is given by

It can be seen that the two waves are in phase when

and are half a cycle out of phase when


3

Constructive interference occurs when the waves are in phase, and destructive interference when
they are half a cycle out of phase. Thus, an interference fringe pattern is produced, where the
separation of the maxima is

and df is known as the fringe spacing. The fringe spacing increases with increase in wavelength,
and with decreasing angle θ.

The fringes are observed wherever the two waves overlap and the fringe spacing is uniform
throughout.

 Between two spherical waves

A point source produces a spherical wave. If the light from two point sources overlaps, the
interference pattern maps out the way in which the phase difference between the two waves varies
in space. This depends on the wavelength and on the separation of the point sources. The figure to
the right shows interference between two spherical waves. The wavelength increases from top to
bottom, and the distance between the sources increases from left to right.

When the plane of observation is far enough away, the fringe pattern will be a series of almost
straight lines, since the waves will then be almost planar.

 Multiple beams

Interference occurs when several waves are added together provided that the phase differences
between them remain constant over the observation time.

It is sometimes desirable for several waves of the same frequency and amplitude to sum to zero
(that is, interfere destructively, cancel). This is the principle behind, for example, 3-phase power
and the diffraction grating. In both of these cases, the result is achieved by uniform spacing of the
phases.
4

It is easy to see that a set of waves will cancel if they have the same amplitude and their phases
are spaced equally in angle. Using phasors each wave can be represented as for waves
from to , where

To show that

one merely assumes the converse, then multiplies both sides by

The Fabry–Pérot interferometer uses interference between multiple reflections.

A diffraction grating can be considered to be a multiple-beam interferometer, since the peaks


which it produces are generated by interference between the light transmitted by each of the
elements in the grating. Feynman suggests that when there are only a few sources, say two, we call
it "interference", as in Young's double slit experiment, but with a large number of sources, the
process is labelled "diffraction".

1.3 Optical interference

Because the frequency of light waves (~1014 Hz) is too high to be detected by currently available
detectors, it is possible to observe only the intensity of an optical interference pattern. The intensity
of the light at a given point is proportional to the square of the average amplitude of the wave. This
can be expressed mathematically as follows. The displacement of the two waves at a point r is:

where A represents the magnitude of the displacement, φ represents the phase and ω represents
the angular frequency.

The displacement of the summed waves is


5

The intensity of the light at r is given by

This can be expressed in terms of the intensities of the individual waves as

Thus, the interference pattern maps out the difference in phase between the two waves, with
maxima occurring when the phase difference is a multiple of 2π. If the two beams are of equal
intensity, the maxima are four times as bright as the individual beams, and the minima have zero
intensity.

The two waves must have the same polarization to give rise to interference fringes since it is not
possible for waves of different polarizations to cancel one another out or add together. Instead,
when waves of different polarization are added together, they give rise to a wave of a
different polarization state.

The discussion above assumes that the waves which interfere with one another are monochromatic,
i.e. have a single frequency—this requires that they are infinite in time. This is not, however, either
practical or necessary. Two identical waves of finite duration whose frequency is fixed over that
period will give rise to an interference pattern while they overlap. Two identical waves which
consist of a narrow spectrum of frequency waves of finite duration, will give a series of fringe
patterns of slightly differing spacings, and provided the spread of spacings is significantly less than
the average fringe spacing, a fringe pattern will again be observed during the time when the two
waves overlap.

Conventional light sources emit waves of differing frequencies and at different times from
different points in the source. If the light is split into two waves and then re-combined, each
individual light wave may generate an interference pattern with its other half, but the individual
fringe patterns generated will have different phases and spacings, and normally no overall fringe
pattern will be observable. However, single-element light sources, such as sodium- or mercury-
vapor lamps have emission lines with quite narrow frequency spectra. When these are spatially
and colour filtered, and then split into two waves, they can be superimposed to generate
6

interference fringes. All interferometry prior to the invention of the laser was done using such
sources and had a wide range of successful applications.

A laser beam generally approximates much more closely to a monochromatic source, and it is
much more straightforward to generate interference fringes using a laser. The ease with which
interference fringes can be observed with a laser beam can sometimes cause problems in that stray
reflections may give spurious interference fringes which can result in errors.

Normally, a single laser beam is used in interferometry, though interference has been observed
using two independent lasers whose frequencies were sufficiently matched to satisfy the phase
requirements.

It is also possible to observe interference fringes using white light. A white light fringe pattern can
be considered to be made up of a 'spectrum' of fringe patterns each of slightly different spacing. If
all the fringe patterns are in phase in the centre, then the fringes will increase in size as the
wavelength decreases and the summed intensity will show three to four fringes of varying colour.
Young describes this very elegantly in his discussion of two slit interference. Some fine examples
of white light fringes can be seen here. Since white light fringes are obtained only when the two
waves have travelled equal distances from the light source, they can be very useful in
interferometry, as they allow the zero path difference fringe to be identified.

Fig 1.2 White light interference fringes and colored interference pattern in a bubble
7

1.4 Principle of interferometry

Fig 1.3 . The light path through a Michelson interferometer

Interferometry refers to a family of techniques in which waves,


usually electromagnetic, are superimposed in order to extract information about the
waves. Interferometry is an important investigative technique in the fields of astronomy, fiber
optics, engineering metrology, optical metrology, oceanography, seismology, spectroscopy (and
its applications to chemistry), quantum mechanics, nuclear and particle physics, plasma
physics, remote sensing, biomolecular interactions, surface profiling, microfluidics, mechanical
stress/strain measurement, and velocimetry Interferometers are widely used in science and industry
for the measurement of small displacements, refractive index changes and surface irregularities.
In analytical science, interferometers are used in continuous wave Fourier transform
spectroscopy to analyse light containing features of absorption or emission associated with a
substance or mixture. An astronomical interferometer consists of two or more separate telescopes
that combine their signals, offering a resolution equivalent to that of a telescope of diameter equal
to the largest separation between its individual elements.
8

1.5 Basic concepts of interferometry

Fig 1.4. Formation of fringes in a Michelson interferometer

Fig 1.5. Colored and monochromatic fringes in a Michelson interferometer:


9

(a) White light fringes where the two beams differ in the number of phase inversions; (b) White
light fringes where the two beams have experienced the same number of phase inversions;

(c) Fringe pattern using monochromatic light (sodium D lines)

Interferometry makes use of the principle of superposition to combine waves


in a way that will cause the result of their combination to have some meaningful property that is
diagnostic of the original state of the waves. This works because when two waves with the
same frequency combine, the resulting pattern is determined by the phase difference between the
two waves—waves that are in phase will undergo constructive interference while waves that are
out of phase will undergo destructive interference. Most interferometers use light or some other
form of electromagnetic wave.

Typically, a single incoming beam of coherent light will be split into two identical beams by
a beam splitter (a partially reflecting mirror). Each of these beams travels a different route, called
a path, and are recombined before arriving at a detector. The path difference, the difference in the
distance traveled by each beam, creates a phase difference between them. It is this introduced
phase difference that creates the interference pattern between the initially identical waves. If a
single beam has been split along two paths, then the phase difference is diagnostic of anything that
changes the phase along the paths. This could be a physical change in thepath length itself or a
change in the refractive index along the path.

As seen in Fig. 1.4 a and b, the observer has a direct view of mirror M1 seen through the beam
splitter, and sees a reflected image M'2 of mirror M2. The fringes can be interpreted as the result of
interference between light coming from the two virtual images S'1 and S'2 of the original source S.
The characteristics of the interference pattern depend on the nature of the light source and the
precise orientation of the mirrors and beam splitter. In Fig. a, the optical elements are oriented so
that S'1 and S'2 are in line with the observer, and the resulting interference pattern consists of circles
centered on the normal to M1 and M'2. If, as in Fig. b, M1 and M'2 are tilted with respect to each
other, the interference fringes will generally take the shape of conic sections (hyperbolas), but
if M1 and M'2 overlap, the fringes near the axis will be straight, parallel, and equally spaced. If S
10

is an extended source rather than a point source as illustrated, the fringes of Fig. a must be observed
with a telescope set at infinity, while the fringes of Fig. 1.4 b will be localized on the mirrors.

Use of white light will result in a pattern of colored fringes. The central fringe representing equal
path length may be light or dark depending on the number of phase inversions experienced by the
two beams as they traverse the optical system

1.6 Categories of interferometric techniques

Interferometers and interferometric techniques may be categorized by a variety of criteria:

1. Homodyne versus heterodyne detection

In homodyne detection, the interference occurs between two beams at the same wavelength (or
carrier frequency). The phase difference between the two beams results in a change in the intensity
of the light on the detector. The resulting intensity of the light after mixing of these two beams is
measured, or the pattern of interference fringes is viewed or recorded. The heterodyne technique
is used for

(1) shifting an input signal into a new frequency range as well as

(2) amplifying a weak input signal (assuming use of an active mixer).

A weak input signal of frequency f1 is mixed with a strong reference frequency f2 from a local
oscillator (LO). The nonlinear combination of the input signals creates two new signals, one at the
sum f1 + f2 of the two frequencies, and the other at the difference f1 − f2. These new frequencies
are called heterodynes. Typically only one of the new frequencies is desired, and the other signal
is filtered out of the output of the mixer. The output signal will have an intensity proportional to
the product of the amplitudes of the input signals.
The most important and widely used application of the heterodyne technique is in
the superheterodyne receiver (superhet), invented by U.S. engineer Edwin Howard Armstrong in
1918. In this circuit, the incoming radio frequency signal from the antenna is mixed with a signal
from a LO and converted by the heterodyne technique to a lower fixed frequency signal called
11

theintermediate frequency (IF). This IF is amplified and filtered, before being applied to
a detector which extracts the audio signal, which is sent to the loudspeaker.
Optical heterodyne detection is an extension of the heterodyne technique to higher (visible)
frequencies.

2. Double path versus common path

A double path interferometer is one in which the reference beam and sample beam travel
along divergent paths. Examples include the Michelson interferometer, the Twyman-Green
interferometer, and the Mach-Zehnder interferometer. After being perturbed by interaction with
the sample under test, the sample beam is recombined with the reference beam to create an
interference pattern which can then be interpreted.

A common path interferometer is a class of interferometer in which the reference beam and
sample beam travel along the same path. Fig. 1.6 illustrates the Sagnac interferometer, the fibre
optic gyroscope, the point diffraction interferometer, and the lateral shearing interferometer. Other
examples of common path interferometer include the Zernike phase contrast microscope, Fresnel's
biprism, the zero-area Sagnac, and the scatterplate interferometer.

Figure 1.6. Four examples of common path interferometers

In a Sagnac interferometer, both beams emerging from the beamsplitter simultaneously go


around an enclosed area in opposite directions and recombine at the original beamsplitter. The
result is an interferometer that is, to first order, completely insensitive to movement of its optical
components, and which offers excellent contrast and extreme fringe stability.
12

The best-known application of the Sagnac interferometer is as a rotation sensor. The first
accounts of the effects of rotation on this form of interferometer were published in 1913
by Georges Sagnac, who mistakenly believed that his ability to detect a "whirling of the ether"
proved the existence of the luminiferous aether, thus disproving special relativity. However, what
he really detected was the independence of the speed of light from the speed of the source, which
is also predicted by special relativity. The sensitivity of present-day Sagnac rotation sensors far
exceeds that of Sagnac's original arrangement. The sensitivity to rotation is proportional to the
area circumscribed by the counter-rotating beams, and fibre optic gyroscopes, present-day
descendants of the Sagnac interferometer, use thousands of loops of optical fibre rather than
mirrors, such that even small to medium sized units easily detect the rotation of the Earth. Ring
laser gyroscopes (not illustrated) are another form of Sagnac rotation sensor that have important
applications in inertial guidance systems.
Besides their use as rotation sensors, interferometers constructed using the Sagnac
topology played an important role in experiments leading to Einstein's discovery of special
relativity, and in the subsequent defense of relativity against theoretical and experimental
challenges. For example, one year before their famous experiment of 1887, Michelson and
Morley (1886) performed a repeat of the Fizeau experiment of 1851, replacing Fizeau's setup
with an even-reflection Sagnac interferometer of such high stability, that even placing a lighted
match in the light path did not cause artifactual fringe displacement.In 1935, Gustaf Wilhelm
Hammar disproved a theoretical challenge to special relativity that attempted to explain away the
null results of Michelson–Morley–type experiments as being a mere artifact of aether-dragging,
using an odd-reflection Sagnac interferometer. He could operate this interferometer in the open,
on a high hilltop with no temperature control, yet still achieve readings of 1/10 fringe accuracy.

Another common path interferometer useful in lens testing and fluid flow diagnostics is
the point diffraction interferometer (PDI), invented by Linnik in 1933. The reference beam is
generated by diffraction from a small pinhole, about half the diameter of the Airy disk, in a
semitransparent plate. Fig. 1.6 illustrates an aberrated wavefront focused onto the pinhole. The
diffracted reference beam and the transmitted test wave interfere to form fringes. The common
path design of the PDI brings to it a number of important advantages.
13

(1) Only a single laser path is required rather than the two paths required by the Mach-Zehnder or
Michelson designs. This advantage can be very important in large interferometric setups such as
in wind tunnels that have long optical paths through turbulent media.

(2) The common path design uses fewer optical components than double path designs, making
alignment much easier, as well as reducing cost, size, and weight, especially for large setups.

(3) While the accuracy of a double path interferometer is dependent on the precision with which
the reference element is figured, careful design enables the generated reference beam of the PDI
to be of guaranteed precision.

A disadvantage is that the amount of light getting through the pinhole depends on how well the
light can be focused onto the pinhole. If the incident wavefront is severely aberrated, very little
light may get through. The PDI has seen use in various adaptive optics applications.

Lateral shearing interferometry is a self-referencing method of wavefront sensing. Instead


of comparing a wavefront with a reference wavefront, lateral shearing interferometry interferes a
wavefront with a shifted version of itself. As a result, it is sensitive to the slope of a wavefront, not
the wavefront shape per se. The illustrated plane parallel plate interferometer has unequal path
lengths for the test and reference beams. Because of this, it must be used with highly
monochromatic (laser) light. The surfaces normally do not have any reflective coating so as to
minimize ghost reflections. An aberrated wavefront from a lens under test is reflected from the
front and back of the plate to form the interference pattern. Variations on this basic design allow
testing of mirrors. Other forms of lateral shearing interferometer, based on
the Jamin, Michelson, Mach–Zehnder, and other interferometer designs, have compensated paths
and may be used with white light.Besides optical testing, applications of lateral shearing
interferometry have included thin film analysis, refractive index measurement, collimation testing,
and adaptive optics.Shearing interferometers, a general framework which includes the lateral
shearing, Hartmann, Shack-Hartmann, rotational shearing, folding shearing, and aperture
masking interferometers, are used in most of the wavefront sensors industrially developed.
14

3. Wavefront splitting versus amplitude splitting

A wavefront splitting interferometer divides a light wavefront emerging from a point or a


narrow slit (i.e. spatially coherent light) and, after allowing the two parts of the wavefront to travel
through different paths, allows them to recombine. Fig. 1.7 illustrates Young's interference
experiment and Lloyd's mirror. Other examples of wavefront splitting interferometer include the
Fresnel biprism, the Billet Bi-Lens, and the Rayleigh interferometer.

Figure 1.7 Two wavefront splitting interferometers

In 1803, Young's interference experiment played a major role in the general acceptance of
the wave theory of light. If white light is used in Young's experiment, the result is a white central
band of constructive interferencecorresponding to equal path length from the two slits, surrounded
by a symmetrical pattern of colored fringes of diminishing intensity. In addition to continuous
electromagnetic radiation, Young's experiment has been performed with individual
photons,withelectrons,and with buckyball molecules large enough to be seen under an electron
microscope.

Lloyd's mirror generates interference fringes by combining direct light from a source (blue
lines) and light from the source's reflected image (red lines) from a mirror held at grazing
incidence. The result is an asymmetrical pattern of fringes. Interestingly, the band of equal path
length, nearest the mirror, is dark rather than bright. In 1834, Humphrey Lloyd interpreted this
effect as proof that the phase of a front-surface reflected beam is inverted.

An amplitude splitting interferometer uses a partial reflector to divide the amplitude of the
incident wave into separate beams which are separated and recombined. Fig. 1.8 illustrates
the Fizeau, Mach–Zehnder and Fabry–Pérot interferometers. Other examples of amplitude
15

splitting interferometer include the Michelson, Twyman–Green, Laser Unequal Path, and Linnik
interferometer.

Figure 1.8 . Three amplitude-splitting interferometers: Fizeau, Mach–Zehnder, and Fabry


Perot

The Fizeau interferometer is shown as it might be set up to test an optical flat. A precisely
figured reference flat is placed on top of the flat being tested, separated by narrow spacers. The
reference flat is slightly beveled (only a fraction of a degree of beveling is necessary) to prevent
the rear surface of the flat from producing interference fringes. Separating the test and reference
flats allows the two flats to be tilted with respect to each other. By adjusting the tilt, which adds a
controlled phase gradient to the fringe pattern, one can control the spacing and direction of the
fringes, so that one may obtain an easily interpreted series of nearly parallel fringes rather than a
complex swirl of contour lines. Separating the plates, however, necessitates that the illuminating
light be collimated. Fig 1.8 shows a collimated beam of monochromatic light illuminating the two
flats and a beam splitter allowing the fringes to be viewed on-axis.

The Mach–Zehnder interferometer is a more versatile instrument than the Michelson


interferometer. Each of the well separated light paths is traversed only once, and the fringes can
be adjusted so that they are in focus in any desired plane.For a wind tunnel study, as illustrated in
Fig. 1.8, the fringes would customarily be adjusted to lie in the same plane as the test object, so
that fringes and test object can be photographed together. If it is decided to produce fringes in
white light, then, since white light has a limited coherence length, on the order of microns, great
care must be taken to equalize the optical paths or no fringes will be visible. A compensating cell
would be placed in the path of the reference beam to match the test cell. Note also the precise
orientation of the beam splitters. The reflecting surfaces of the beam splitters would be oriented so
16

that the test and reference beams pass through an equal amount of glass. In this orientation, the
test and reference beams each experience two front-surface reflections, resulting in the same
number of phase inversions. The result is that light traveling an equal optical path length in the test
and reference beams produces a white light fringe of constructive interference.

The heart of the Fabry–Pérot interferometer is a pair of partially silvered glass optical flats
spaced several millimeters to centimeters apart with the silvered surfaces facing each other.
(Alternatively, a Fabry–Pérot etalon uses a transparent plate with two parallel reflecting
surfaces.)As with the Fizeau interferometer, the flats are slightly beveled. In a typical system,
illumination is provided by a diffuse source set at the focal plane of a collimating lens. A focusing
lens produces what would be an inverted image of the source if the paired flats were not
present; i.e. in the absence of the paired flats, all light emitted from point A passing through the
optical system would be focused at point A'. In Fig. 1.8, only one ray emitted from point A on the
source is traced. As the ray passes through the paired flats, it is multiply reflected to produce
multiple transmitted rays which are collected by the focusing lens and brought to point A' on the
screen. The complete interference pattern takes the appearance of a set of concentric rings. The
sharpness of the rings depends on the reflectivity of the flats. If the reflectivity is high, resulting in
a high Q factor (i.e. high finesse), monochromatic light produces a set of narrow bright rings
against a dark background. In Fig. 1.8, the low-finesse image corresponds to a reflectivity of 0.04
(i.e. unsilvered surfaces) versus a reflectivity of 0.95 for the high-finesse image.

It is interesting to note that Michelson and Morley (1887)and other early experimentalists
using interferometric techniques in an attempt to measure the properties of the luminiferousaether,
used monochromatic light only for initially setting up their equipment, always switching to white
light for the actual measurements. The reason is that measurements were recorded visually.
Monochromatic light would result in a uniform fringe pattern. Lacking modern means
of environmental temperature control, experimentalists struggled with continual fringe drift even
though the interferometer might be set up in a basement. Since the fringes would occasionally
disappear due to vibrations by passing horse traffic, distant thunderstorms and the like, it would
be easy for an observer to "get lost" when the fringes returned to visibility. The advantages of
white light, which produced a distinctive colored fringe pattern, far outweighed the difficulties of
17

aligning the apparatus due to its low coherence length. This was an early example of the use of
white light to resolve the “2 pi ambiguity”

Fig 1.9 Modern Michelson interferometer setup


18

1.7 Perspex:
Perspex, Acrylite, and Lucite are all names for the same polymer commonly classified under
‘glass’ for economic and commercial reasons. It is often called Acrylic glass. Chemically, it is
the synthetic polymer of Methyl methacrylate abbreviated as PMMA. It is usually used as an
alternative to Polycarbonate when such extreme levels of strength are unnecessary. Additionally,
PMMA does not contain the potentially harmful bisphenol-A subunits found in polycarbonate. It
is often preferred because of its moderate properties, easy handling and processing, and low cost.
Non-modified PMMA behaves in a brittle manner when loaded, especially under an impact
force, and is more prone to scratching than conventional inorganic glass, but modified PMMA
can achieve high scratch and impact resistance.

Fig 1.10 Graph showing the variation of refractive index of Perspex with wavelength of light used
19

Interlude: Illustration of Michelson interferometer setup for


common laboratory use

The setup shows an air-pump and air-chamber as well, usually intended for the determination of
change in refractive index of a gas in response to applied pressure (usually air). This project
however makes no use of the Pressure variation apparatus and instead it is replaced by a glass slab
holder with a micrometer screw.
20

Chapter 2

Determination of Thickness of Glass Slab using Michelson


Interferometer (Using He-Ne laser source).

2.1 Introduction

The aim of this project work centers around calculating the thickness of a glass slab using
Michelson Interferometer (Using a He-Ne laser source).

Here we use a shock resistant board, a diode laser, a laser mount, a beam splitter mount, mirror
mount (2 numbers), Screen & Rotation stage with Glass mounted for the experimental setup.

2.2 Theory

In principle, the method for calculating the index of refraction is relatively simple. The
light passes through a greater length of glass as the plate is rotated. The general steps for measuring
the index of refraction in such a case are as follows:

1. Determine the change in the path length of the light beam as the glass plate is rotated.
Determine how much of the change in path length is through glass, dg, and how much is
through air, da.

2. Relate the change in path length to the measured fringe transitions with the following
equation:

𝜃
2𝑛𝑎 𝑑𝑎 (𝜃) + 2𝑛𝑔 𝑑𝑔
𝜆0
21

Where na= the index of refraction of air, ng = the index of refraction of the glass Plate (given for a
specific wavelength), λ0= the wavelength of the light source in vacuum, and N= the number of
fringe transitions that has been counted.

If we assume identical optical constants for air and vacuum as is often done in the case of a first
order approximation. Hence we let λ0 = λair . As the glass plate is rotated, more and more material
is in front of the light beam and hence the optical path length varies ever so slightly with the angle
turned. This can be easily worked out, either using Calculus or following a small angle
approximation. Whichever way the calculations are made, the thickness is given by

𝑁λair {𝑛 − (1 − cos 𝜃)}


𝑡=
2(𝑛 − 1)(1 − cos 𝜃)

Where t is the thickness of the glass slab, n the refractive index of the material of the glass slab at
the wavelength used (here n being 1.515 for glass). ‘θ’ the angle through which the slab is turned.
N being the number of fringes sunken.

2.3 Experimental procedure

The following procedures are to be done

1 Align the laser and interferometer in the Michelson mode.


2 Place the rotation stage between the beam-splitter and movable mirror, perpendicular to the
optical path.
3 Mount the glass plate on the rotation stage.
4 Position the stage and glass such that degree is 0 and a glass slide is perpendicular to the
optical path.
5 When glass plate introduced in the optical path of Michelson interferometer, the fringe will
be shifted & will become blur. To make the fringe sharpen again, move the Mirror mount
to & fro till the clear set of fringes is achieved on the viewing screen.
6 Slowly rotate the rotation stage. Count the number of fringe transitions that occur as one
rotates the table from 0 degrees to an angle θ (at least 10 degrees)
22

Calibrating the Micrometer:-

For even more accurate measurements of the mirror movement, one can use a laser to
calibrate the micrometer.

To do this, set up the interferometer in Michelson mode. Turn the micrometer knob
you count off at least 20 fringes. Carefully note the change in the micrometer reading,
and record this value as d’. The actual mirror movement, d., is equal to N λ/2, where
λ is the known wavelength of the light (0.6350 μm for the Diode Laser) and N is the
number of fringes that were counted. In future measurements, multiply your
micrometer readings by d/d’ for a more accurate measurement.

2.4 Experimental Observation:

To find the thickness of the glass plate using screw gauge

P.S.R. (mm) Observed H.S.R. Corrected H.S.R. Thickness t (mm)


2 95 93 2.93
2 95 93 2.93
2 94 92 2.92
2 96 94 2.94
2 95 93 2.93

Mean t = 2.93 mm
23

To find the angle of rotation for sinking of N fringes:

Initial Heading Final Heading Angle turned (θ) Number of fringes


sunken (N)

90 98 8° 30

96 104 8° 30

100 108 8° 30

270 278 8° 30

260 268 8° 30

280 288 8° 30

282 290 8° 30

300 307 7° 30

80 88 8° 30

82 90 8° 30

Mean θ = 7.9

Wavelength of laser beam in air , 𝜆 air=635 x10-9 m

Thickness of the glass plate using Michelson’s interferometer is given by

𝑁λair {𝑛 − (1 − cos 𝜃)}


𝑡=
2(𝑛 − 1)(1 − cos 𝜃)

30 ∗ 635 ∗ 10−9 {1.515 − (1 − cos 7.9)}


𝑡=
2(1.515 − 1)(1 − cos 7.9)
24

𝑡 = 0.002934 𝑚

𝑡 = 2.934 ∗ 10−3 𝑚 = 2.934 𝑚𝑚

Percentage of deviation from the measurement from the screw gauge is given by

|𝑒1 − 𝑒2 |
𝛥= ∗ 100
𝑒1

2.934 − 2.930
𝛥= ∗ 100
2.934

𝛥 = 0.136%
25

Chapter 3

Determination of Refractive index of a Perspex using Michelson


Interferometer (Using He-Ne laser source).

3.1 Introduction

The aim of this project work also centers around calculating the Index of Refraction of
Perspex using Michelson Interferometer (Using a He-Ne laser source).

Here we use a shock resistant board, a diode laser, a laser mount, a beam splitter mount, mirror
mount (2 numbers), Screen & Rotation stage with Perspex mounted for the experimental setup.

3.2 Theory

In principle, the method for calculating the index of refraction is relatively simple. The
light passes through a greater length of Perspex as the plate is rotated. The general steps for
measuring the index of refraction in such a case are as follows:

3. Determine the change in the path length of the light beam as the Perspex plate is rotated.
Determine how much of the change in path length is through Perspex, dp, and how much
is through air, da.

4. Relate the change in path length to the measured fringe transitions with the following
equation:

𝜃
2𝑛𝑎 𝑑𝑎 (𝜃) + 2𝑛𝑝 𝑑𝑝
𝜆0
26

Where na= the index of refraction of air, np = the index of refraction of the Perspex Plate (as yet
unknown), λ0= the wavelength of the light source in vacuum, and N= the number of fringe
transitions that has been counted.

If we assume identical optical constants for air and vacuum as is often done in the case of a first
order approximation. Hence we let λ0 = λair . As the Perspex plate is rotated, more and more material
is in front of the light beam and hence the optical path length varies ever so slightly with the angle
turned. This can be easily worked out, either using Calculus or following a small angle
approximation. Whichever way the calculations are made, the thickness is given by

(2𝑡 − 𝑁𝜆𝑎𝑖𝑟 )(1 − cos 𝜃)


𝑛=
2𝑡(1 − cos 𝜃) − 𝑁𝜆𝑎𝑖𝑟

Where t is the thicknes of the Perspex slab, n the refractive index of the material of the Perspex
slab at the wavelength used (here n being unknown). ‘θ’ the angle through which the slab is turned.
N being the number of fringes sunken.

3.3 Experimental procedure

The following procedures are to be done

7 Align the laser and interferometer in the Michelson mode.


8 Place the rotation stage between the beam-splitter and movable mirror, perpendicular to the
optical path.
9 Mount the Perspex plate on the rotation stage.
10 Position the stage and Perspex such that degree is 0 and a Perspex slide is perpendicular to
the optical path.
11 When Perspex plate introduced in the optical path of Michelson interferometer, the
fringe will be shifted & will become blur. To make the fringe sharpen again, move
the Mirror mount to & fro till the clear set of fringes is achieved on the viewing screen.
12 Slowly rotate the rotation stage. Count the number of fringe transitions that occur as one
rotates the table from 0 degrees to an angle θ (at least 10 degrees)
27

Calibrating the Micrometer:-

For even more accurate measurements of the mirror movement, one can use a laser to
calibrate the micrometer.

To do this, set up the interferometer in Michelson mode. Turn the micrometer knob
you count off at least 20 fringes. Carefully note the change in the micrometer reading,
and record this value as d’. The actual mirror movement, d., is equal to N λ/2, where
λ is the known wavelength of the light (0.6350 μm for the Diode Laser) and N is the
number of fringes that were counted. In future measurements, multiply your
micrometer readings by d/d’ for a more accurate measurement.

3.4 Experimental Observation:

To find the thickness of the Perspex plate using screw gauge

P.S.R. (mm) Observed H.S.R. Corrected H.S.R. Thickness t (mm)


2 88 88 2.88
2 89 89 2.89
2 89 89 2.89
2 90 90 2.90
2 88 88 2.88

Mean t = 2.888 mm
28

To find the angle of rotation for sinking of N fringes:

Initial Heading Final Heading Angle turned (θ) Number of fringes


sunken (N)

300 308 8° 30

280 288 8° 30

270 278 8° 30

210 219 9° 30

90 98 8° 30

80 88 8° 30

82 90 8° 30

100 108 8° 30

110 118 8° 30

112 120 8° 30

Mean θ = 8.1

Wavelength of laser beam in air , 𝜆 air=635 x10-9 m

Index of refraction of Perspex

(2𝑡 − 𝑁𝜆𝑎𝑖𝑟 )(1 − cos 𝜃)


𝑛=
2𝑡(1 − cos 𝜃) − 𝑁𝜆𝑎𝑖𝑟

(2 ∗ 2.888 ∗ 10−3 − 30 ∗ 635 ∗ 10−9 )(1 − cos 8.1)


𝑛=
2 ∗ 2.888(1 − cos 8.1) − 30 ∗ 635 ∗ 10−9

𝑛 = 1.4887
29

CONCLUSION

In the present project work I have carried out an extensive study on interferometry by
analyzing various literature on the subject. An experimental study on Michelson Interferometry
was also carried out and the thickness of a given sample of glass slab was calculated within 0.136%
of the screw gauge reading (the interferometer reading being 2.934mm). Using the same setup, the
refractive index of Perspex glass or more commonly known as Acrylic glass was also found to be
1.4887. Both were conducted using a Helium-Neon Laser source of 635nm wavelength.
30

REFERENCES

1. Optics 10th Edition – Eugene Hecht

2. Optics 3rd Edition – Frank L Pedrotti, Leno M Pedrotti, Leno S Pedrotti

3. Refractiveindex.info (For graphing and data resources)

4. Wave Optics – R K Verma

5. Advanced Optics & Microwave Laboratory Manual Edited by V.S.Jayakumar, K Joy,

Jijimon K Thomas and Hubert Joe, Department of Physics, Mar Ivanios College.

6. Principles of Optics – Max Born, Emil Wolf

7. Interferometry – W H Steel
8. Basics of Interferometry – P Hariharan

9. The Feynman Lectures in Physics (Vol 1) The definitive edition – Richard P Feynman

10. Vibrations and Waves – A P French

11. Fundamentals of Physics 10th edition – Jearl Walker, David Halliday and Robert Resnick

12. University Physics 13th edition – Sears and Zemansky

You might also like