Lecture 6 (2023)

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Lecture 6

6.1 Optical fibers


Optical fibers use total internal reflection (TIR) to transmit light signals and images. A
typical optical fiber is made of a high refractive index core and a low refractive index coating
(cladding), as shown in Figure 1. Some fibers are provided with EMA (extra mural
absorption), to suppress evanescent waves, which are waves “leaking” from the core, despite
the TIR phenomenon. Light needs to be incident approximately parallel to the fiber axis, so
that it can strike the walls at angles higher than the critical angle of the core/cladding interface
and be thus totally reflected. As a result, light is trapped inside; once it is input, it will
continue to reflect almost without loss off the walls of the fiber. Optical fibers (single or
bundles) are the main components of fiber optic cables which are commonly used in
telecommunications.

Figure 1. Optical path of light in an optical fiber

As we have already mentioned, for the light to be confined inside the fiber, the angle
of incidence to the core/cladding interface must be greater than the limiting angle for total
internal reflection. It has already been demonstrated that the critical angle of incidence θ c
corresponds to a refraction angle of 90°.

Figure 2. The structure of an optical fiber.

In figure 2, the refractive index of the core is denoted by n1 and the refractive index of the
cladding is n2 . According to Snell’s law of refraction, at the point where light strikes the fiber
wall,
n2
sin θ c = . (0.1)
n1
To undergo total internal reflection light should be input at an angle less than a maximum
value α corresponding to θ c . That value of sin α is called the numeric aperture NA of the
optical fiber.
The numeric aperture is easily computed as a function of the refractive indices of the
optical fiber. From Snell’s law and (0.1):
( )
NA= sin α = n1 sin 90 − θ c = n1 cos θ c =

n22 (0.2)
= n1 1 − sin 2 θ c = n1 1 − = n12 - n22
n12
If n2= n1 − ∆n , where ∆n is a small quantity, the last member above can be rewritten as

NA
= (
n12 − n12 − 2∆n ⋅ n1 + ( ∆n )
2
)≈ 2n1 ⋅ ∆n , (0.3)

because ( ∆n )  2n1 ⋅ ∆n . The value of NA gives the input angle range which is necessary
2

for a ray of light to be transmitted in the fiber.


The main applications of fiber optics are in the field of medicine and communications.
In medicine, optical fibers are used to visualize various internal organs. The fibers are tiny,
flexible, have little loss and allow for straightforward construction of the images. All these
qualities make them extremely suitable for medical investigations. In communications, optical
fibers are used to transmit information at a much higher rate than systems using radio
frequencies, since the rate of transmission is related to the signal frequency. Also, the volume
of information is huge: a single fiber the thickness of a hair could transmit audio information
equivalent to 32000 voices speaking simultaneously. If we add to that the fact that optical
fibers are not subject to electrical interference and are long-life devices, you will understand
why they are preferred to the classic telecommunication systems. Original patents on fiber
optic transmission were granted to John Loggie Baird in the 1930's.

6.2 Geometrical and optical path length


Let us consider that light propagates between two points P and Q. The geometrical
path is the actual distance traveled by light starting from P and ending at Q, denoted by l .
When light propagates in a homogeneous medium, characterized by a constant refractive
index n , the optical path length is defined as the product between the geometrical path length
l and the refractive index of the medium n , and is denoted by [l ] :
[l ] = nl (0.4)
The optical path length in vacuum is equal to the geometrical path, [lvac ] = l . In any other
medium the optical path length is longer [l ] > [lvac ] , since n > 1 .
Light travels at its highest speed c in a vacuum. In a transparent medium, it is slowed
down and it travels at a speed which is n times smaller than c . Say that light takes ∆t
seconds to travel a distance equal to the geometrical path l , in a medium whose refractive
index is n . In the same time interval, light would travel a distance n times longer, if it were
propagating in vacuum. Indeed, we have
l l nl [l ]
∆t = = = = (0.5)
v cn c c
So, the optical path length [l ] is the distance that light would travel in vacuum, in the
time necessary for it to travel a geometrical path l in a medium whose optical properties
are described by the refractive index n . Obviously, the higher the refractive index, the
smaller the speed is, and the optical path length is consequently longer.
Note that the concept of optical path is designed to consider both the actual distance
traveled by light and its speed. It is a crucial concept for understanding light propagation in a
dispersive medium (a medium whose refractive index depends on wavelength). Since any
transparent medium is more or less dispersive, we understand that optical path length is an
extremely useful tool for any piece of optics.
The optical path length determines the propagation time, the difference in phase
introduced by light traveling a geometrical path l :
2π 2π 2π 2π
δ= k ∆= l l
⋅ ∆= l
⋅ ∆= ⋅ n∆=
l ⋅ [ ∆l=
] kvac [ ∆l ] (0.6)
λ λvac n λvac λvac

and the number N of wavelengths that span a geometrical path l :

N= =
l l
=
nl
=
[l ] (0.7)
λ λvac n λvac λvac
Note that the propagation of light in a medium can be described in terms of propagation in
vacuum, simply by taking the precaution of replacing the geometrical paths by the
corresponding optical paths. The concept of optical path can be used to compare paths of light
through arbitrary, optically different media.

6.3 Fermat’s Principle


Fermat’s principle originally stated that of all the geometrically possible paths that
light could take between two points P and Q in a medium, the actual path is that of the least
time of propagation. In other words, the propagation time has a minimum value on the
real path. According to equation (0.5), the time of propagation is determined by the optical
path length. So, Fermat’s principle requires that the optical path on the real path is minimum.
The modern formulation of this principle is that of all the geometrically possible paths that
light could take between two arbitrary points P and Q in a medium, the actual path is that
which corresponds to a stationary value of the optical path. By “stationary” we mean a
minimum or maximum value This modern formulation is more general than the initial one.
If an optically transparent medium is nonhomogeneous, the refractive index is a
function of position, n = n ( l ) . Then the optical path length is correctly defined as
[l ] = ∫ n(l )dl , (0.8)
C

where C is path of light, meaning the curve that connects the arbitrary points P and Q.
Fermat’s principle states that the real path Cactual corresponds to a stationary value of the
optical path length integral, that is
d [l ]
=0 (0.9)
dl C =C
actual

Note that once found, the actual optical path length stays the same for any pair of
arbitrary points. Hence, if light propagates from P to Q on Cactual , it will take the same path
from Q to P. This assertion is known as the principle of reversibility of light rays. Now we
can see that it results as a direct consequence of Fermat’s principle.
In the following, we will apply Fermat’s principle to derive the laws of reflection and
refraction. The demonstrations are
P simple and elegant, avoiding
cumbersome geometrical
Q constructions. Let us consider that
light gets from P to Q after perfect
reflection by a mirror (surface Σ ) -
p see Figure 3. There are an infinity
q geometrically possible paths which
θi θ r meet this condition. However, path
POQ is the actual one. Light travels
P1 Q1 only in air, whose refractive index is
Σ
almost unity. The total optical path is
O the sum of the optical path from P to a
x d−x point O on the plane of the mirror and
the optical path from O to Q.
d

Figure 3. The path of light from P to Q after reflection on Σ , at point O.

With the notations given in Figure 3, we have


[l ]= p2 + x2 + q2 + ( d − x ) ,
2
(0.10)

[l ] n=
where = air l l is the overall optical path length, since nair ≅ 1 . Please note that x is a
variable: it can take in principle any value, and the length of the optical path depends on it,
[l ] = function ( x ) . According to Fermat's principle, on the real path of light between P and Q,
the derivative of the optical path length [l ] with respect to x must be zero, for the condition
of stationarity to be met.
d [l ]
= 0. (0.11)
dx real position of O

Let us denote by xreal the correct position of point O.

d [l ] 1 1
( ) 
−1/2
( )
−1/2
⋅ 2 ( d − x ) q2 + ( d − x )
2
=  ⋅ 2x ⋅ p2 + x2 −  = 0. (0.12)
dx x = xreal
2 2  x = xreal

After calculation, we obtain that


xreal ( d − xreal )
= , (0.13)
p2 + x2 q 2 + ( d − xreal )
2

Now we go back to Figure 3 and notice that in the triangles PP1O and QQ1O the cosine
functions of the complementary angles of θi and θ r can be expressed as
OP1 xreal
( )
cos 90 − θi = sin θi =
OP
=
2
, (0.14)
p 2 + xreal
OQ1 ( d − xreal )
( )
cos 90 − θ r = sin θ r =
OQ
= , (0.15)
q 2 + ( d − xreal )
2

Since θi and θ r belong to the range 0,90  it results that

θ r = θi , (0.16)

which says that light travels from P to Q on the real path for which the angle of reflection is
equal to the angle of incidence, and this is what the law of reflection states.
To get a better insight of Fermat’s principle, we will demonstrate the law of refraction
as well. In this case, P and Q lie in two different dielectric transparent media separated by the
surface Σ (see Figure 6), which are described by the indices of

θi
ni
P1 O Q1
Σ
nt
x q
θt
d

Figure 4. The path of light from P to Q after refraction on Σ , at point O.

refraction ni and nt . The total optical path is the sum of the optical path [l ]i of the ray in the
medium where light is incident and the optical path [l ]t in the medium where light is
refracted.
[l ] = [l ]i + [l ]t = ni p 2 + x 2 + nt q 2 + ( d - x ) ,
2
(0.17)

First, we calculate the derivative:


d [l ] d [l ]i d [l ]t
= + =
dx dx dx
ni
=
d
dx
( )
p 2 + x 2 + nt
d  2
dx 
2 
 q + (d - x)  =

, (0.18)

x (d - x)
= ni − nt ,
p2 + x2 q2 + ( d - x )
2

then we apply Fermat’s principle (0.11),

d [l ] xreai ( d - xreal )
ni
= − nt 0,
=
dx x = xreal
2 2
p + xreal q 2 + ( d - xreal )
2

which results in the following equation


ni sin θi − nt sin θt =
0, (0.19)
This is Snell’s law of refraction. As expected, by applying Fermat’s principle we obtained that
the actual path of light that crosses the boundary surface between two dielectric media is
given by the law of refraction.

6.4 Huygens Principle. Huygens-Fresnel Principle


In 1670 Christiaan Huygens used the wave theory to account for the laws of geometric
optics. He suggested a way of understanding the propagation of waves by postulating that as a
wave propagates through a medium, each point on the advancing wave front acts as a new
point source of the wave and formulated a principle which is known as Huygens’ Principle. It
states that ”every point on a primary wave front serves as a source of secondary spherical
wavelets, which advance with a speed v equal to that of the primary wave front at each
point in space; the primary wave front at some later time is the envelope of these wavelets”.
Basically, this is a geometrical technique for finding the path of a wave, regardless its
wavelength. All the points on the primary wave front are thus considered sources of
secondary wavelets spreading out in all directions, with a speed equal to the wave speed in the
medium. One can draw arcs of a circle with a radius vt , then draw the surface tangent, or
envelope, to find the new wave front at a later time t (see Figure 5).
Figure 5. Huygens construction of the wavefronts at different time moments as envelopes of the
secondary wavelets (2D representation)
According to Huygens' Principle, the light rays associated with a plane light wave in
empty space propagate in straight lines. It is straightforward to account for the laws of
reflection and refraction using Huygens' Principle. However, the principle fails to explain why,
for example, an expanding spherical wave continues to expand outward from its source, rather
than converging back toward the source. In other words, Huygens overlooks the fact that the
wave front formed by the back half of the wavelets implies a light disturbance traveling in the
opposite direction. Also, it does not account for diffraction of the light into the region of
geometric shadow when light is incident on obstacles or apertures.
Later, Augustin Fresnel, an adept of Maxwell’s electromagnetic theory, amended
Huygens’ Principle to eliminate its above-mentioned flaws. The result is known as the
Huygens-Fresnel Principle which states: Every unobstructed point of a wave front is a
source of spherical secondary wavelets with the same frequency as that of the primary. The
amplitude of the resultant wave at any forward point is the superposition of these wavelets,
considering their amplitudes and relative phase. In other words, the secondary waves
mutually interfere, and this accounts for light diffraction, as we are going to show in the
following chapters.

P
 
n r

α
dS

Figure 6. Huygens-Fresnel Geometry. S is an arbitrary wavefront; dS is an elementary surface whose


   
normal in n ; r is the position vector of the point of observation P; α is the angle between r and n .

According to Huygens-Fresnel Principle, the equation which expresses the electric


field intensity in the spherical wavelet emitted by an arbitrary elementary area dS , belonging
to an arbitrary wave front S , at a point of observation P, located at the distance r from dS ,
at any time t , can be written as
E dS
= dE ( r,t ) const (α ) 0 cos (ω t − kr + ϕ0 ) , (0.20)
r

where E0 and φ 0 are the amplitude and initial phase of the electric field oscillation, ω and k
are the angular frequency and the wave number of the wavelets (Figure 4). The quantity
const (α ) is a constant of proportionality which depends on the angle α between the

direction of the normal n to the elementary surface dS and the direction of the vector

position r of the observation point. It describes how waves propagate at maximum intensity
in the forward direction and cannot propagate back to the source. Thus
const ( 0 ) = 1 , (0.21)
And
( 0,
const α ≥ 900 = ) (0.22)

In the equation which gives dE ( r , t ) , the 1 r factor is due to the stated spherical character of
the wavelets. Thus, their amplitudes diminish even when the waves propagate in a
nonabsorbent medium.
The resultant wave at P is obtained by superposing all the secondary wavelets
generated by the wave front S:
E dS
= EP = ∫ dE ∫ const (α ) 0 cos (ω t − kr + ϕ0 ) , (0.23)
S S
r

This equation is an analytic expression of the Huygens-Fresnel Principle. It is used to


compute the intensity of the diffracted light at any point in space and any time. The
calculation of the integral is however far from being easy. It presumes knowing the angle α
for any infinitesimal surface dS belonging to the primary front wave S, at any point P in
space. For an arbitrary surface, the calculations involve sophisticated approximating
techniques. Fortunately, where highly symmetric front surfaces are concerned, the calculation
reduces to a mere algebraic summation, as we are going to see in the lectures dedicated to the
diffraction phenomenon.

You might also like