You are on page 1of 170

Condition for spatial coherence

The degree of spatial coherence of a beam of light can be deduced from the
contrast of the fringes produced by it. The broader the source of the light, the
lesser is the degree of coherence. In Young’s double slit experiment, it the slits S1
and S2 are directly illuminated by a source, interference fringes are not observed.
Instead the screen is uniformly illuminated. The absence of fringes indicates that
the light issuing from the slits do not possess spatial coherence. If a narrow slit is
introduced before the double slit, light passing through the narrow slit S
illuminates S1 and S2. The waves emerging from them, having been derived
through wave front division, are coherent and stationary interference pattern will
be observed on the screen. If the width of the slit S is gradually increased the
contrast of fringe pattern decreases and fringes disappear. When the slit S is
wider, S1 and S2 receive waves from different part of S which do not maintain
coherence. When S is narrow, it ensures that the wave trains incident on slits S1
and S2 originate from a small region of the source and hence they have spatial
coherence.

Formation of distinct fringe pattern depends on two parameters in the double slit
experiment. One is the size of the slit S and second is the separation d between
the two slits S1 and S2.
From Figure it is seen that the path difference between the two waves passing at
the edges A and B of the slit S is

Δ = b sinα

Distinct fringes will be obtained when Δ«λ/2.

Therefore, it requires that b sinα « λ/2.

This equation determines the size of the source, which is spatially coherent to
produce fringes of satisfactory contrast. Above equation shows that the smaller is
the size of the source, the better is the spatial coherence. Spatial coherence of
waves decreases with increasing size of the light source.
Fresnel’s Biprism
Fresnel used a biprism to show interference phenomenon. The biprism abc
consists of two acute angled prisms placed base to base. Actually, it is constructed
as a single prism of obtuse angle of about 179o. The acute angle α on both sides is
about 30’. The prism is placed with its refracting edge parallel to the line source S
(slit) such that Sa is normal to the face bc of the prism. When light falls from S on
the lower portion of the biprism it is bent upwards and appears to come from the
virtual source B. Similarly light falling from S on the upper portion of the prism is
bent downwards and appears to come from the virtual source A. Therefore A and
B act as two coherent sources. Suppose the distance between A and B = d. If a
screen is placed at C, interference fringes of equal width are produced between E
and F but beyond E and F fringes of large width are produced which are due to
diffraction. To observe the fringes, the screen can be replaced by an eye-piece
and fringes are seen in the field of view. If the point C is at the principal focus of
the eye piece, the fringes are observed in the field of view.
Question: In Fresnel’s biprism arrangement shows that the distance between two
virtual coherent sources is 2y1(μ – 1)α, where y1 is the distance between source
and biprism, α is the angle of biprism and μ is the refractive index of the material
of the biprism.

Answer: For a prism of very small refracting angle α, the deviation δ produced in
ray by the prism is given by

δ = (μ – 1)α

where μ is the refractive index of the material of the prism.

From figure it is clear that the distance d between the sources S1 and S2 is given
by,

d = 2δ×y1

where y1 is the distance between the slit and biprism.

Finally, we get

d = 2(μ-1)αy1
Fringes with White Light Using a Biprism
When white light is used, the centre of the fringe at C is white while the fringes on both sides of
C are coloured because the fringe width β depends upon wavelength. In a biprism, the two
coherent virtual sources are produced by refraction and the distance between the two sources
depends upon the refractive index, which in turn depends upon the wavelength of light.
Therefore, for blue light the distance between the two apparent sources is different to that
with red light. The distance of the nth fringe from the centre (with monochromatic light)

Therefore for blue and red rays, the nth fringe will be,
Expression for Fringe Width (Similar to double slit experiment)
Let the distance between S1 and S2, the two virtual sources be d and D is the distance of the
screen from the sources on which the fringes are obtained. The point C on the screen is
equidistant from S1 and S2. Therefore, the waves from S1 and S2 reach C in the same phase
and reinforce each other. So the point C will be the centre of bright fringe while the illumination
(brightness or darkness) at any other point P on the screen will depend upon the path
difference between the waves from S1 and S2.
Change of Phase Due to Reflection (Lloyd’s Mirror)

Young’s method for producing two coherent light sources involves illuminating a
pair of slits with a single source. Another simple, yet ingenious, arrangement for
producing an interference pattern with a single light source is known as Lloyd’s
mirror. A point light source is placed at point S close to a mirror, and a viewing
screen is positioned some distance away and perpendicular to the mirror.

Light waves can reach point P on the screen either directly from S to P or by the
path involving reflection from the mirror. The reflected ray can be treated as a ray
originating from a virtual source at point S’. As a result, we can think of this
arrangement as a double-slit source with the distance between points S and S’
comparable to length d in Figure. Hence, at observation points far from the
source (L » d) we expect waves from points S and S’ to form an interference
pattern just like the one we see from two real coherent sources. An interference
pattern is indeed observed. However, the positions of the dark and bright fringes
are reversed relative to the pattern created by two real coherent sources (Young’s
experiment). This can only occur if the coherent sources at points S and S’ differ in
phase by 180°.
To illustrate this further, consider point P’, the point where the mirror intersects
the screen. This point is equidistant from points S and S’. If path difference alone
were responsible for the phase difference, we would see a bright fringe at point P’
(because the path difference is zero for this point), corresponding to the central
bright fringe of the two-slit interference pattern. Instead, we observe a dark
fringe at point P’. From this, we conclude that a 180° phase change must be
produced by reflection from the mirror. In general, an electromagnetic wave
undergoes a phase change of 180° upon reflection from a medium that has a
higher index of refraction than the one in which the wave is traveling.
Interference due to reflected light
Interference due to transmitted light
Determination of Wavelength of Light
Fringes of Equal Inclination
Michelson Interferometer

A schematic diagram of the Michelson interferometer is shown in Fig. 1; S represents a light


source (which may be a sodium lamp) and L represents a ground glass plate so that an
extended source of almost uniform intensity is formed. G1 is a beam splitter; i.e., a beam
incident on G1 gets partially reflected and partially transmitted. M1 and M2 are good-quality
plane mirrors having very high reflectivity. One of the mirrors (usually M2) is fixed and the
other (usually M1) is capable of moving away from or toward the glass plate G1 along an
accurately machined track by means of a screw. In the normal adjustment of the
interferometer, mirrors M1 and M2 are perpendicular to each other and G1 is at 45o to the
mirror. Waves emanating from a point P get partially reflected and partially transmitted by the
beam splitter G1, and the two resulting beams are made to interfere in the following manner:
The reflected wave (shown as 1 in figure) undergoes a further reflection at M1, and this
reflected wave gets (partially) transmitted through G1; this is shown as 5 in the figure. The
transmitted wave (shown as 2 in figure) gets reflected by M2 and gets (partially) reflected by G1
and results in the wave shown as 6 in the figure. Finally, waves 5 and 6 interfere and give
interference pattern. This can be easily seen from the fact that if x1 and x2 are the distances of
mirrors M1 and M2 from the plate G1, then to the eye the waves emanating from point P will
appear to get reflected by two parallel mirrors (M1 and M2’) separated by a distance x1 ~ x2.
And, if we use an extended source, then no definite interference pattern will be obtained on a
photographic plate placed at the position of the eye. Instead, if we have a camera focused for
infinity, then on the focal plane we will obtain circular fringes, each circle corresponding to a
definite value of θ.

In an actual Michelson interferometer, the beam splitter G1 consists of a plate (which may be
about 1-2 cm thick), the back surface of which is partially silvered, and the reflections occur at
the back surface as shown in Fig. 2. It is immediately obvious that beam 5 traverses the glass
plate three times, and to compensate for this additional path, one introduces a “compensating
plate” G2 which is exactly of the same thickness as G1.
Now, if the beam splitter is just a simple glass plate, the beam reflected from mirror M2 will
undergo an abrupt phase change of π (when getting reflected by the beam splitter), and since
the extra path that one of the beams will traverse will be 2(x1 ~ x2), the condition for
destructive interference will be

2d cos θ = mλ, where m = 0, 1, 2, 3, . . . . and d = x1 ~ x2

and the angle θ represents the angle that the rays make with the axis which is normal to the
mirrors. Similarly, the condition for a bright ring is

2d cos θ = (2m+1)λ/2.
Measurement of wavelength
Determination of the difference in the wavelength of two
waves
Multiple Beam Interference
When the interference pattern is obtained due to the superposition of many coherent beams
together then it is known as multiple beam interference. For example, if a plane wave falls on a
plane parallel glass plate, then the beam undergoes multiple reflections at the two surfaces and
a large number of beams of successively diminishing amplitude emerge on both sides of the
plate. These beams (on either side) interfere to produce an interference pattern at infinity. The
fringes formed are much sharper than those by two-beam interference and, therefore, the
interferometers involving multiple-beam interference have a high resolving power and hence
find applications in high resolution spectroscopy.
Fabry-Perot Interferometer
Intensity Distribution
When the interference pattern is obtained due to the superposition of many coherent beams
together then it is known as multiple beam interference. For example, if a plane wave falls on a
plane parallel glass plate, then the beam undergoes multiple reflections at the two surfaces and
a large number of beams of successively diminishing amplitude emerge on both sides of the
plate. These beams (on either side) interfere to produce an interference pattern at infinity. The
fringes formed are much sharper than those by two-beam interference and, therefore, the
interferometers involving multiple-beam interference have a high resolving power and hence
find applications in high resolution spectroscopy.
Anti Reflection Coatings
Dielectric Mirrors
Interference Filters
Diffraction Grating
An arrangement which essentially consists of a large number of equidistant slits is known as
a diffraction grating; the corresponding diffraction pattern is known as the grating spectrum.
Since the exact positions of the principal maxima in the diffraction pattern depend on the
wavelength, the principal maxima corresponding to different spectral lines (associated with a
source) will correspond to different angles of diffraction. Thus the grating spectrum provides
us with an easily obtainable experimental setup for determination of wavelengths. For
narrow principal maxima (i.e., sharper spectral lines), a large value of N is required (I α N2).
A good-quality grating, therefore, requires a large number of slits (typically about 15,000 per
inch). This is achieved by ruling grooves with a diamond point on an optically transparent
sheet of material; the grooves act as opaque spaces. After each groove is ruled, the machine
lifts the diamond point and moves the sheet forward for the ruling of the next groove.
Rowland’s arrangement gave 14,438 lines per inch, corresponding to d = 2.54/14,438 =
1.759 × 10–4 cm. For such a grating, for λ = 6 × 10–5 cm, the maximum value of m would be
2, and, therefore, only the first two orders of the spectrum will be observed. However, for λ =
5 × 10–5 cm, the third-order spectrum will also be visible.

The positions of the principal maxima are given by

d sin θ = mλ m = 0, 1, 2, . . . . (1)

This relation, which is also called the grating equation, can be used to study the dependence
of the angle of diffraction θ on the wavelength λ. The zeroth-order principal maximum occurs
at θ = 0 irrespective of the wavelength. Thus, if we are using a polychromatic source (e.g.,
white light), then the central maximum will be of the same color as the source itself.
However, for m ≠ 0, the angles of diffraction are different for different wavelengths, and
therefore, various spectral components appear at different positions. Thus by measuring the
angles of diffraction for various colors one can (knowing the value of m) determine the
values of the wavelengths. The intensity is maximum for the zeroth-order spectrum (where
no
dispersion occurs), and it falls off as the value of m increases.

If we differentiate Eq. (1), we obtain

Δθ/Δλ = m/d cosθ (Dispersive Power) (2)

Assuming θ to be very small (i.e., cosθ ≈ 1), we can see that the angle Δθ is directly
proportional to the order of spectrum m for a given Δλ, so that for a given m, Δθ/Δλ is a
constant. Such a spectrum is known as a normal spectrum, and in this the difference in angle
for two spectral lines is directly proportional to the difference in wavelengths.
Resolving Power of Grating:

The resolving power refers to the power of distinguishing two nearby spectral lines and is
defined by the

R = λ/Δλ

where Δλ is the separation of two wavelengths which the grating can just resolve; the smaller
the value of Δλ, the larger the resolving power.

The Rayleigh criterion can be used to define the limit of resolution. According to this
criterion, if the principal maximum corresponding to the wavelength λ+ Δλ falls on the first
minimum (on the either side of the principal maximum) of the wavelength λ, then the two
wavelengths λ and λ + Δλ are said to be just resolved.

If this common diffraction angle is represented by θ and if we are looking at the mth-order
spectrum, then the two wavelengths λ and λ + Δλ will be just resolved if the following two
equations are simultaneously satisfied:

d sin θ = m (λ + Δλ) and

d sin θ = mλ + λ/N

Thus

R = λ/Δλ = mN

which implies that the resolving power depends on the total number of lines in the grating.
Further, the resolving power is proportional to the order of the spectrum. Thus to resolve the
D1 and D2 lines of sodium (Δλ= 6 Å) in the first order, N must be at least (5.89 × 10–5)/(6 ×
10–8) = 1,000.
Fresnel Half Period Zones
Diffraction by a Circular Aperture
Zone Plate
Comparison between a zone plate and a convex lens

1. Both the zone plate and convex lens form a real image of the object and the
equations connecting the conjugate distances are similar.
2. The focal lengths of both depend on the wavelength and hence suffer from
chromatic aberration.
3. In case of zone plate the image is formed by the diffraction phenomenon.
In case of a convex lens the image is formed due to refraction of light.
4. The zone plate has got multiple foci on either side of the plate. Hence the
intensity of the image formed will be much less. Convex lens has only one
focus. As all the light is focused at one point the intensity of the image will
be more.
5. In a zone plate waves reaching the image point through any two alternate
zones differ in path by λ and in phase by 2π. In case of a convex lens all the
rays reaching the image point have zero path or phase difference.
6. A zone plate can be used over a wide range of wavelengths from
microwave to x-rays. Glass lens cannot be used beyond the visible region.
Half Period Strips
(Intensity due to Cylindrical Wavefront)
Diffraction by a Straight Edge
Diffraction Due to a Narrow Wire
Polarization
Types of Polarization
Representation of Polarized and Unpolarized Light
Brewster’s Law


Double Refraction and Birefringence
Physics Behind Double Refraction and Birefringence

Birefringence is the optical property of a material having a refractive index that depends on
the polarization and propagation direction of light. These optically anisotropic materials are
said to be birefringent (or birefractive). The birefringence is often quantified as the maximum
difference between refractive indices exhibited by the material.
Birefringence is responsible for the phenomenon of double refraction whereby a ray of light,
when incident upon a birefringent material, is split by polarization into two rays taking slightly
different paths.
Uniaxial material
The simplest type of birefringence is described as uniaxial, meaning that there is a single
direction governing the optical anisotropy whereas all directions perpendicular to it (or at a
given angle to it) are optically equivalent. Thus rotating the material around this axis does not
change its optical behavior. This special direction is known as the optic axis of the material.
Light whose polarization is perpendicular to the optic axis is governed by a refractive
index no (for "ordinary"). Light whose polarization is in the direction of the optic axis sees an
optical index ne (for "extraordinary"). For any ray direction there is a linear polarization
direction perpendicular to the optic axis, and this is called an ordinary ray. However, for ray
directions not parallel to the optic axis, the polarization direction perpendicular to the ordinary
ray's polarization will be partly in the direction of the optic axis, and this is called
an extraordinary ray. The ordinary ray will always experience a refractive index of no, whereas
the refractive index of the extraordinary ray will be in between no and ne, depending on the ray
direction. The magnitude of the difference is quantified by the birefringence:

Δn = ne - no
The propagation (as well as reflection coefficient) of the ordinary ray is simply described
by no as if there were no birefringence involved. However the extraordinary ray, as its name
suggests, propagates unlike any wave in a homogenous optical material. Its refraction (and
reflection) at a surface can be understood using the effective refractive index (a value in
between no and ne).
When the light propagates either along or orthogonal to the optic axis, such a lateral shift does
not occur. In the first case, both polarizations see the same effective refractive index, so there is
no extraordinary ray. In the second case the extraordinary ray propagates at a different phase
velocity (corresponding to ne) but is not an inhomogeneous wave.
Although there is no axis of symmetry, there are two optical axes or binormals which are
defined as directions along which light may propagate without birefringence, i.e., directions
along which the wavelength is independent of polarization. For this reason, birefringent
materials with three distinct refractive indices are called biaxial. Additionally, there are two
distinct axes known as optical ray axes or biradials along which the group velocity of the light is
independent of polarization.
Double refraction
When an arbitrary beam of light strikes the surface of a birefringent material, the polarizations
corresponding to the ordinary and extraordinary rays generally take somewhat different
paths. Unpolarized light consists of equal amounts of energy in any two orthogonal
polarizations, and even polarized light will have some energy in each of these polarizations.
According to Snell's law of refraction, the angle of refraction will be governed by the
effective refractive index which is different between these two polarizations. The different
angles of refraction for the two polarization components are shown in the figure at the top of
the page, with the optic axis along the surface (and perpendicular to the plane of incidence), so
that the angle of refraction is different for the p polarization (the "ordinary ray" in this case,
having its polarization perpendicular to the optic axis) and the s polarization (the "extraordinary
ray" with a polarization component along the optic axis).
Huygen’s Theory of Double Refraction
in a Uniaxial Crystal
Production of Elliptically and Circularly Polarised Light
from Linearly Polarised Light
Or
Plane Polarised Light and Circularly Polarised Light are
Special Cases of Elliptically Polarized Light
Quarter-Wave Plate
Half-Wave Plate
Calcite Crystal and Nicol Prism
Production and Detection of Polarised Light
Optical Rotation
Physics Behind Optical Activity (Optical Rotation)
Laurent’s Half-Shade Polarimeter
Bi-Quartz Polarimeter
Babinet Compensator

You might also like