Optics I 2017 Chapter2 PDF

Download as pdf or txt
Download as pdf or txt
You are on page 1of 44

E.

Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

Chapter 2

Propagation of light

In free space (vacuum), light propagates with constant phase velocity c. Most real situations,
however, involve propagation in matter. In that case, the perpetual interaction between the
electromagnetic wave and the atoms, molecules, or solid constituting the medium of
propagation, modifies the characteristics of the propagating wave. In particular, when the
propagating light is of frequency that matches resonant absorption due to electronic
transitions in the medium, strong attenuation of the electromagnetic field occurs.
In this chapter, however, we will mainly be concerned with light of non-resonant
frequency that is different from that of electronic transitions in the medium. In this case,
weak attenuation of the amplitude of the electromagnetic field results and, more importantly,
the phase velocity changes. In addition to the propagation in a homogeneous medium with
such non-resonant interaction, we will also formulate the phenomena of reflection and
refraction at boundaries between different homogeneous media. The case of geometrical
optics as a limiting case of light propagation will also be discussed.

2.1 Light Scattering


We consider the propagation of an electromagnetic field of frequency ω in matter consisting
of atoms whose electronic transitions are non-resonant with this frequency. That is, the
photon energy is different from that required for inducing transitions between atomic or
molecular energy levels characterizing the medium. This results in elastic scattering (rather
than absorption) of the incident photons impinging on the atom or molecule in question. In
the simplest picture, the incident plane wave will give rise to a scattered spherical wave, as
illustrated in Fig 2.1.
To evaluate the impact of this scattering on the incident wave, we use a simple,
classical scattering model. We will show that this scattering leads to a modification in the
phase velocity of the wave, establishing a particular refractive index in the medium.

Figure 2.1: A plane electromagnetic wave scattered off an atom.

1
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

2.1.1 Scattering Model


To model the scattering process, consider a plane wave of frequency ω incident on a single
“scattering center” representing an atom in the medium. The atom is treated as a harmonic
oscillator of a characteristic frequency ω 0 ≠ ω . The impinging electromagnetic field induces
in the atom an electric dipole

p (t ) = p 0 cos(ωt )

This electric dipole radiates with a characteristic electromagnetic field (in vacuum) given by

p0 ω 2 cos ( kr − ω t ) E
E= 2
sin (θ ) ; B=
4πε 0 c r c

where θ is the polar angle. The radiation pattern of such a dipole is illustrated schematically
in Fig. 2.2.

Figure 2.2: Schematic illustration of the radiation pattern of an electric dipole.

In the radiation zone, far away from the dipole, the electric and magnetic fields are
perpendicular to each other, and the time-averaged Poynting vector becomes:

ε 0c 2 p02ω 4 sin 2 θ
I= S = E0 =
T
2 32π 2ε0c 3 r 2

Notice that the intensity varies as the inverse of the fourth power of the wavelength (
I ~ λ−4 ~ ω 4 ).

To cast a relationship between the amplitudes of the incident and the scattered wave,
we use a classical, damped harmonic oscillator model to describe the atom. The displacement
x (t ) of the electron (charge -e), viewed as a mass m attached to the nucleus of the atom by a
spring, then obeys the equation of motion

2
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

2 F −eE
x˙˙ + σx˙ + ω0 x = =
m m

where ω 0 is the non-damped frequency of the oscillator, σ is the damping constant and E is
the electric field providing the driving force. We treat the electromagnetic wave in the scalar
€ iωt iωt
approximation. For a monochromatic field of the form E(t ) = E0 e we expect x (t ) = x0 e ,
and by inserting into the equation of motion we get

−eE0
−ω 2 x0 + iωσx0 + ω02 x0 =
m

Solving for the amplitude of oscillation, we obtain


€ −eE0 1
x0 =
m ω0 − ω 2 + iσω
2

This complex amplitude can be written as x0 = x0 e − iα , with magnitude



eE 0 /m eE 0 /m
x0 = 1
  → 1/ 2
ω ≈ ω0
[(ω − ω ) + σ ω ]
2
0
2 2 2 2 2
[(2ω 0Δω) 2
+ σ 2ω 02 ]
and phase
€  σω   σ 
α = tan−1 2 2
  → tan−1 
ω0 − ω  ω ≈ ω0  2Δω 

From these expressions for the response represented by the complex amplitude x0, we see that
the largest response in amplitude occurs at resonance ( ω = ω 0 ); this response has a

Lorentzian lineshape (i.e., dependence on frequency) and its peak decreases with increasing
damping σ. The frequency dependence of the amplitude and phase of the oscillator are
depicted in Fig. 2.3.

Figure 2.3: Frequency dependence of the amplitude (left) and phase (right) of a harmonic
damped oscillator. The solid line is for smaller damping, dashed for larger damping.

We now use this simple model to derive an expression for the polarization of N such
oscillators (per unit volume). Starting with the definition:

3
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

P = −Nex0 = εχ E

we obtain

Ne 2 E 0 (ω ) 1
P (ω ) =
m ω 0 − ω 2 + iσω
2

Using the relation


€  ε(ω ) 
P (ω ) = ε0 χ (ω ) E 0 (ω ) = ε0  −1 E 0 (ω )
 ε0 

we also get the susceptibility


€ Ne 2 1 ε (ω )
χ (ω ) = 2 2 = −1
ε 0 m ω 0 − ω + iσω ε0

Writing the complex dielectric constant in terms of the complex refractive index n˜ (ω ) as

ε(ω )
= n˜ 2 (ω )
ε0 €
and using our result
€ε ω
( ) = 1+ Ne 2 1
ε0 ε0 m ω − ω 2 + iσω
2
0

we obtain

Ne2 ω02 − ω 2 − iσω
n˜ 2 (ω ) = 1+
ε 0 m (ω02 − ω 2 )2 + σ 2ω 2

The real and imaginary parts of the refractive index, defined as


€ n˜ 2 (ω ) = nr2 (ω ) + ini2 (ω )

are then given by

€  Ne 2 ω 02 − ω 2
2
 n r (ω ) = 1+
 ε0 m (ω 02 − ω 2 ) 2 + σ 2ω 2
 2
 n 2 ω = − Ne σω
 i ( )
ε0 m (ω 02 − ω 2 ) 2 + σ 2ω 2


4
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

The real part of the refractive index (depicted in Fig. 2.4) yields the correction to the phase
velocity in the medium filled with the given atoms/molecules, whereas the imaginary part
determines the attenuation constant.

Figure 2.4: Frequency dependence of the real part of the refractive index near the
resonant frequency ω 0 .

The simple model described in this section provides the two important elements

characterizing the scattering of an electromagnetic wave by a scattering center: the spatial
distribution of the scattered wave and the frequency response of the medium. Next, we use
these characteristics to describe the impact on the propagating wave in two important limiting
cases, the dilute scattering medium and the dense scattering medium (see Fig. 2.5). In a dilute
medium, the distance between scattering centers is much larger than the wavelength, whereas
it is smaller than the wavelength in a dense medium.

Figure 2.5: Schematic illustration of a plane wave impinging on a dilute (left) and a dense
(right) scattering medium. The scattering centers are represented by filled circles.

2.1.2 Dilute Scattering Centers (Rayleigh Scattering)


In this case, referred to as Rayleigh scattering, the average distance between the scattering
centers in the medium of propagation is much larger than the wavelength. Each scattering
center gives rise to a scattered wavelet whose phase is independent of the other scattered
wavelets. The combined effect of all the scattering centers on the incident beam depends on
the interference of all the scattered wavelets. The result of this interference, in turn, depends
strongly on the scattering direction with respect to the direction of the incident wave.

5
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

Lateral scattering:

Here, the distance from the scattering center to the point of observation varies considerably
for the different wavelets. Hence, the random phases of the wavelets lead to an incoherent,
−4
weak lateral scattering, with scattered intensity that varies as λ . This Rayleigh scattering
mechanism explains, for example, the blue color of sun light scattered by particles in the
−4
atmosphere. For larger scatterers, the distinct λ wavelength dependence disappears; the
resulting wavelength-independent scattering, so-called Mie scattering, is responsible, for
example, for the white color of milk.

Forward scattering:

In this situation the number of wavelengths within the distance to the observation point does
not vary appreciably from one scattering center to another, and the scattered wavelets sum up
in a coherent way. The constructive interference in the forward direction leads to the forward
propagation of the beam.

2.1.3 Dense Scattering Centers (Coherent Scattering)


If the average distance between the scattering centers is much smaller than the wavelength,
the nature of the interference, which gives rise to the lateral and forward scattering, changes
considerably. This case corresponds, for instance, to the propagation in a homogeneous solid
medium such as glass, where the scattering centers are the atoms/molecules themselves.

Lateral scattering:

Since there are many scattering centers within a distance of one wavelength, many scattered
wavelets result with all possible phases, and the interference is completely destructive. There
is thus no laterally scattered wave in this case.

Forward scattering:

The forward propagating wave is the result of the coherent superposition of the scattered
wavelets. The various interfering wavelets differ in their phase. In fact, the delay in the
secondary (scattered) wave with respect to the primary (incident) one gives rise to a smaller
phase velocity, which is reflected in the value of the refractive index being larger than one
(see Fig. 2.6).
Note that the frequencies of the incident wave and the resulting forward wave are the
same ( ν 0 ). The smaller phase velocity of the propagating wave can be written then

c
V p < c ⇒ Vp =

n

with n>1, and the wavelength in the medium becomes:


€ V p (c /n ) (c /ν 0 ) λ0
λ= = = =
ν0 ν0 n n

6 €
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

where λ 0 = c / ν 0 is the wavelength in vacuum.


The discussion above shows that the amplitude of the laterally scattered wave usually
decreases with increasing density in the propagation medium. In homogeneous solids and
liquids, the lateral scattering is usually weak, and is mainly determined by the density of
scattering centers that have separations larger than the wavelength, e.g., defects and density
fluctuations.

Figure 2.6: Contributions of the primary and scattered waves to the sum of the
electromagnetic field amplitude, for two different secondary waves.

2.2 Huygens’ and Fermat’s Principles


Although the propagation of an optical wave with an arbitrary wavefront profile can be
calculated directly from Maxwell’s equation (using Kirchoff’s diffraction theory, see Chapter
6 in Optics I), it is extremely useful to use simplified geometrical principles to predict the
propagation course. Huygens’s and Fermat’s principles give such descriptions that are
successfully applicable in many important situations.

2.2.1 Huygens’s principle


Consider the propagation of a scalar wave in a homogeneous, isotropic medium. Given the
wavefront of this wave at a given instance, Huygens’s principle provides a geometrical
construction procedure for determining the way this wavefront evolves. This geometrical
construction is illustrated in Fig. 2.7.
Huygens’s principle states that:

7
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

Every point on a propagating wavefront serves as the source of a spherical secondary


wavelet such that the wavefront at a later time corresponds to the envelope of these
wavelets.

Figure 2.7: Application of Huygens’ principle for constructing a phase front starting from
a given one. The reconstructed phase front is the envelope of the spherical wavelets whose
origins lie on the original one.

Note that Huygens’s wavelets have the same frequency and phase velocity as those of
the primary, incident wave. Moreover, the direction of propagation of the reconstructed
wavefront is the same as that of the incident one. This conjecture is formally justified by the
more fundamental Kirchoff’s diffraction theory (see Chapter 6 of Optics I). An extension of
this principle to the case of non-isotropic media will be presented in the discussion on
birefringence, in Chapter 4 of Optics I.
As an example, we will now apply Huygens’s principle to derive the laws of
reflection and refraction of a plane wave incident on a planar boundary between two different
homogeneous media characterized by refractive indices n i and n t ≠ n i .

Reflection
€ €
In this case the incident and reflected plane waves, both of angular frequency ω , propagate in
the same medium of index ni. Thus, their wavelengths are the same: λi = λr = (2πc /ω ) /n i .
Using the Huygens construction procedure, we see that the reflected wave has its planar
wavefronts inclined by the same angle as those of the incident€one with respect to the
interface (see Fig. 2.8). We thus obtain the law of reflection for a plane wave impinging on a
planar surface: €

θi = θr

8
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

Figure 2.8: Application of Huygens’s construction to the reflection of a plane wave off a
planar surface.

Refraction

We now consider the wave transmitted to the other side of the interface. In this case the
wavelengths λi = (2πc /ω ) /n i and λt = (2πc /ω ) /n t of the incident and transmitted plane
waves are different. Hence, the propagation directions of the incident and transmitted waves
are in general different, a process referred to as refraction. Using the Huygens construction as
sketched in Fig. 2.9 we get the relation
€ €
λi λt
=
sin θ i sin θ t

where θ i and θ t are the angles for the incident and transmitted plane waves. Inserting the
values of the wavelengths in the two media we obtain

c ni c nt
=
sin θi sinθ t

Figure 2.9: Application of Huygens’s construction to the refraction of a plane wave at a


planar interface between two dielectric media.

This result is equivalent to the well-known Snell’s law describing the refraction of a light ray
at a planar boundary (see Fig. 2.10):

ni sin θ i = nt sin θ t

9
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

Note that the light rays employed in geometrical optics correspond in the wave formulation to
the propagation directions, i.e., the direction normal to the phase fronts of the propagating
waves.

Figure 2.10: Snell’s law: ray optics representation.

2.2.2 Fermat’s principle


This principle concerns the trajectory of an optical ray propagating from a given point A to
another point B in space. In particular, it formulates the “path of choice” for such ray in terms
of the optical path length l, defined as:

B
l= ∫ n( s)ds
A

where s is the curvilinear abscise measured along the trajectory. The time t it takes for a light
ray to propagate along a distance ds is related to the optical path through the relation

ds 1 l
t=∫ = ∫ dsn ( s) =
c /n(s) c c

Fermat’s principle states that:



A light ray, in going from point A to point B, traverses an optical path whose length is
stationary with respect to variations to that path.

In Fig. 2.11(a), the “path of choice” S is indicated by the solid curve connecting the
points A and B. The optical path length corresponding to S is stationary in the sense that
several other “similar” paths (the dashed lines almost coincident with S, in Fig. 2.11(a)) will
have almost the same optical path as that of S. Often the length of the optical path exhibits a
minimum with respect to small variations of the physical path; hence the path of choice
corresponds often to the least time for propagating between the two points.

10
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

Figure 2.11: Fermat’s principle: (a) A path S whose optical path length is stationary with
respect to small variations; (b) application to the problem of refraction at the boundary
between two homogeneous media.

As an application of Fermat’s principle, consider the propagation of a light ray from


point A to point B separated by a planar interface between two homogeneous media of
different index of refraction n1 and n 2 (Fig. 2.11(b)). Denoting the distances of A and B from
the interface by h and b, and the distance between the points (parallel to the interface) by a,
we can write the time of propagation as
€ € 2
h2 + x2 b 2 + (a − x )
t= +
v1 v2

where v1 = c /n i and v 2 = c /n 2 are the phase velocities in the two media and x is the (parallel
to the interface) distance between A and the point where the ray crosses the interface. The
rate of change of this €
time with respect to variations in x should vanish according to Fermat’s
principle. This gives the condition
€ €
dt x − (a − x)
= + =0
dx v1 h 2 + x 2 v b 2 + (a − x)2
2

which yields

sin θ1 sin θ 2
=
v1 v2

Expressing the phase velocities in term of the refractive indices, we recover Snell’s law:

n1 sin θ1 = n2 sin θ 2 .

11
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

2.3 Reflection and Refraction


The simple laws of reflection and refraction described in the previous section rely on
Huygens’s and Fermat’s principles, which do not account for the vector nature of the
electromagnetic field. A more accurate, vectorial description of these phenomena, based on
the boundary conditions for the fields involved, leads to the formulation of Fresnel’s
equations.

2.3.1 Fresnel’s Equations


Consider a plane wave incident on a planar surface, defined by the unit normal vector uˆ n ,
that separates two homogeneous dielectric media i and t (see Fig. 2.12). The electric fields of
the incident, reflected and transmitted waves are given, respectively, by:

Ei ( r, t ) = E 0i cos ( ki ⋅ r − ω i t )
Er ( r, t ) = E 0r cos ( k r ⋅ r − ω r t + φr )
Et ( r, t ) = E 0t cos ( k t ⋅ r − ω t t + φt )

Figure 2.12: Geometry for the derivation of Fresnel’s equations.

The boundary conditions for the electric and magnetic fields at the interface are found
using the integral version of Faraday’s law, with a contour C taken across the interface (see
Fig. 2.12):
d
∫ C E ⋅ dl = − dt ∫ S B ⋅ dA
Taking the limit where the width h of this contour vanishes we get:

h → 0 ⇒ E tang,1 − E tang,2 = 0

⇒ E tang,1 = E tang, 2


12
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

That is, the tangential component of the (total) electric field vector is continuous upon
crossing the interface. A similar argument, based on the integral version of Ampere’s law,
leads to the condition that the tangential component of the magnetic field H is also
continuous at the interface.
Since the projection of a vector onto a plane can be represented by the cross vector
product of that vector and the unit vector normal to the plane, we can rewrite the boundary
condition for the electric field in the form

uˆ n × (E i + E r ) = uˆ n × E t

For this relation to hold, the arguments of the electric field functions should have the same
spatial and time dependence at the interface. Thus, we should have

(ki ⋅ r − ω i t ) int erface = (kr ⋅ r − ω rt + φ r ) int erface = ( kt ⋅ r − ω t t + φt ) int erface

Since this relation must be valid for all values of r and t, we obtain the conditions

(1) ω i = ω r = ω t = ω
 (at the interface)
(2) k i ⋅ r = k r ⋅ r + φ r = k t ⋅ r + φ t

The physical significance of the first condition is that the polarization vector, which gives rise
to the radiated reflected and transmitted fields in response to the incident field, oscillates at
€ frequency as the incident one. This condition yields
the same

k i = k r = n iω /c

The second condition gives

(€
ki − kr ) ⋅ r = const (at the interface)

As the vector r lies in the plane of the interface, this equation defines the plane of the
interface, and we must have:

(k i − k r ) // uˆ n
Hence, this difference vector lies in the plane containing k i and k r , and we can therefore
write:

uˆ n × (k i − k r ) = 0
€ €
or
€ ki sinθ i = k r sinθ r

where the angles define the directions of the wave vectors (see Fig. 2.12). Finally, since the
ω
norms of these k vectors are equal, k i = k r = n i , we get:
c

€ 13
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

θi = θ r

This result is identical to that obtained using the Huygens principle in 2.2.1. Note that the
vectors {k i ,k r , uˆ n } are co-planar. The plane containing these vectors is referred to as the
plane of incidence.
The second condition also gives
€ (ki − kt ) ⋅ r = const (at the interface)

Following a similar reasoning as for the incident and reflected waves, we find

(k i − k t ) // uˆ n
which yields
€ uˆ n × (k i − k t ) = 0

and hence
€ ki sinθ i = k t sin θt

Dividing both sides by k0 = ω /c , we get the condition for the transmission angle:

ni sinθ i = nt sinθ t

This result is €
Snell’s law, also derived previously using the Huygens principle in 2.2.1. Note
that the vectors {k i ,k t , uˆ n } are co-planar as well. Thus, the vectors {k i ,k r ,k t , uˆ n } all lie in the
plane of incidence.
Fresnel’s equations relate the amplitude of the electric field of the incident wave to
those of the reflected and transmitted waves. These relations are obtained by applying the
€ conditions to the electric and magnetic field €vectors at the interface of the two
continuity
homogeneous dielectric media. We distinguish two polarization cases.

E perpendicular to plane of incidence

The three-dimensional geometry and the directions of the electric field vectors are illustrated
in Fig. 2.13 for the case where the electric field vectors are perpendicular to the plane of
incidence. The projections of the electric and magnetic field in the plane of incidence are
shown for this case in Fig. 2.14.

14
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

Figure 2.13: Geometry of the reflected and transmitted waves for electric field vectors
perpendicular to the plane of incidence.

Figure 2.14: Projections of the electric and magnetic fields in the plane of incidence for the
case of a “perpendicular electric field”.

The continuity of the field parallel to the interface then implies

E0i + E0r = E0t

The in-plane component of H is also continuous, which yields

15
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

Bi B B
− cosθ i + r cosθ r = − t cosθ t
µi µr µt

Now, for a plane wave, we have

∂B ∂E x ∂B
∇×E=− ⇔ =− y
∂t ∂z ∂t

and using

 E x = E 0 cos(kz − ωt )

 By = B0 cos(kz − ωt )

we get the relation between the field amplitudes:

c
E0 = B
n 0

Using this relation we obtain the continuity conditions:

 ni nt
 (E 0i − E 0 r )cosθ i = E 0t cosθ t
 µi µt
E + E = E
 0i 0r 0t

which yield the expressions for the amplitude reflection coefficient:

ni n
cos θi − t cos θ t
 E 0r 
 = µi
µt
r⊥ ≡ 
 E 0i  ⊥ n i n
cos θi + t cosθ t
µi µt

and the amplitude transmission coefficient:

ni
2 cosθ i
E  µi
t ⊥ ≡  0t  = 1 + r⊥ =
 E 0i ⊥ ni n
cosθ i + t cosθ t
µi µt

E parallel to plane of incidence

The projections of the electric and magnetic fields in the plane of incidence for this case are
shown in Fig. 2.15. Here, the condition for the continuity of the tangential component of E
gives:

E 0i cosθ i − E 0 r cosθ r = E 0t cosθ t

16
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

and the continuity of the tangential component of H reads:

ni n n
E 0i + r E 0 r = t E 0t
µi µr µt

The amplitude reflection and transmission coefficients are thus given by:

nt n
cosθ i − i cos θt
 E 0r 
r// ≡   = µt µi
 E 0i  // n i n
cosθ t + t cosθ i
µi µt
and
n
2 i cos θi
 E 0t  µi
t// ≡   =
 E 0i  // n i n
cos θt + t cos θi
µi µt

Figure 2.15: Projections of the electric and magnetic fields in the plane of incidence for the
case of a “parallel electric field”.

The expressions given above for the reflection and transmission coefficients constitute the
basic Fresnel’s equations.
A special important case concerns non-magnetic media, for which we can set
µi = µt = µ0 . In this case, the amplitude reflection and transmission coefficients become:

17
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

 n i cos θ i − n t cos θ t
 r⊥ = n cos θ + n cos θ
i i t t
Polarization s (TM) 
t = 2n i cosθ i

 n i cosθ t + n t cos θ i
 n t cosθ i − n i cosθ t
 r// = n cosθ + n cosθ
t i i t
Polarization p (TE) 
€ t = 2n i cos θ i
 // n t cos θ i + n i cosθ t

Here, we have introduced the common notation for the perpendicular (s or TM polarization)
and parallel (p or TE) directions of the electric field with respect to the plane of incidence.

Note that the incidence and transmission angles are not independent. In fact, using
Snell’s law,

ni sinθ t
=
nt sin θi

the refractive indices can be eliminated from the expressions for the amplitude reflection and
transmission coefficients, and we finally obtain:

 sin θ t cosθ i − cosθ t sin θ i sin(θ i − θ t )


 r⊥ = =−
 sin θ t cosθ i + cosθ t sin θ i sin(θ t + θ i )
Polarization s (TM) 
t = 2cosθ i sin θ t 2cos θ i sin θ t
⊥ =
 sin θ i cos θ t + cosθ i sin θ t sin(θ i + θ t )

 cosθ i sin θ i − cos θ t sin θ t tan(θ i − θ t )


 r// = =
€  cosθ t sin θ t + sin θ i cosθ i tan(θ i + θ t )
Polarization p (TE) 
t = 2cos θ i sin θ t 2sin θ t cosθ i
=
 // cos θ t sin θ t + cos θ i sin θ i sin(θ i + θ t ) cos(θ i − θ t )

2.3.2 Properties of Reflected and Transmitted Waves


Several important observations
€ can be made by inspection of Fresnel’s equations.

Normal Incidence:

In this case the incident ray is perpendicular the plane defining the interface between the two
media. We then obtain:

θi = 0 ⇒ θ t = 0

and the amplitude coefficients become:


18
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

r = −r = nt − ni
 // ⊥
nt + ni

2n i
t // = t⊥ =
 nt + n i

This leads to the relationship between the reflection and transmission coefficients:

t// + r// = t ⊥ + (−r⊥ ) = 1

Note that this relationship is valid only in the absence of losses (e.g., due to scattering,
absorption) at the interface.
For non-normal incidence, Fresnel’s equations reveal a rich dependence on the angle
of incidence. We classify this dependence according to the relative values of refractive
indices involved.

External Reflection

For such reflection the incident ray arrives in the low-index medium, that is ni<nt. The
geometry and the incident and transmitted rays involved are shown in Fig. 2.16.

Figure 2.16: External reflection: incident, reflected and transmitted rays.

The dependence of the reflection and transmission coefficients on the angle of incidence is
shown in Fig. 2.17. In this case, for a particular incidence angle θ i = θ B , θi + θt = π /2 and
hence r// = 0 . Thus, only the s-polarization component is reflected; hence, the reflected field
corresponding to an incident field with both s- and p-polarization components (non-
polarized) becomes polarized. This occurs at a particular angle of incidence called the
Brewster angle. The value of Brewster’s angle can be calculated with the aid of Snell’s law:

π
ni sinθ i = nt sin − θ i  = n t cos θi
2 

which yields

19
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

tan θ B = n t n
i

As an example, for an air/glass interface, one finds θB = tan−1 1.5 = 56.3 .


Figure 2.17: Angular dependence of the amplitude reflection and transmission coefficients
in the case of air/glass interface (n=1.5) for external reflection.

Internal Reflection

In this case the incident ray arrives in the medium with the higher refractive index, that is
ni>nt (see Fig. 2.18). The angular dependence of the reflection and transmission coefficients
is shown in Fig. 2.19.

Figure 2.18: Internal reflection: incident, reflected and transmitted rays.

20
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

Figure 2.19: Angular dependence of the amplitude reflection and transmission coefficients
in the case of air/glass interface for internal reflection.

From Snell’s law, we now have θt > θi , and for a critical angle of incidence θi = θ c
we obtain θt = π /2 , i.e., the transmitted wave becomes a surface wave. For larger angles of
incidence, there is no real solution for the transmitted angle; this situation is referred to as
total internal reflection. Snell’s law gives directly the value of the critical angle for obtaining
such total reflection:

ni sinθ i = nt

or

θc = sin−1 (n t /n i )

As examples, for a glass/air interface one finds θc = 41.8 , and for a water/air interface one
has θc = 49.8 .
It is important to note that the reflection coefficients can be negative or positive, and
for angles beyond the total internal reflection angle become complex numbers. In general,
they can be written:

r⊥,// = r⊥,// e iφ

The phase-shift angle φ is plotted in Fig. 2.20 as a function of the angle of incidence for
several cases.


21


E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

Figure 2.20: Phase shift angle of reflection coefficient as a function of angle of incidence,
for several geometries.

22
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

For practical applications, one is often interested in the fraction of optical power
reflected or transmitted at an interface between two dielectric media. To calculate these
power reflection and transmission coefficients R and T, we use the expressions for the
intensity,

1c 1
I= S T
= εE 0 2 = cnε0 E 0 2
2n 2

Denoting the cross section of the beam cut by the interface by A (see Fig. 2.21), we get for
the reflectance:

2
I Acosθ r I r  E 0r 
R≡ r = = = r2
I i Acosθ i Ii  E 0i 

and for the transmittance:

2
I A cosθ t I t cosθ t nt cos θt  E 0t  n t cos θt 2
T≡ t = = = t
Ii A cosθ i Ii cosθ i ni cos θi  E 0i  ni cos θi

In the case of no absorption at the interface, conservation of energy implies

Ii A cosθ i = Ir A cosθ r + I t A cosθ t

or:

I r I t cosθ t
1= + = R+ T
Ii Ii cosθ i

For the special case of normal incidence, the expressions for the reflectance and the
transmittance reduce to:
2
 n − ni 
R = R// = R⊥ =  t 
 n t + ni 

2
n t  2n i  4n i n t
T = T// = T⊥ =   =
ni  nt + ni  (nt + ni )2

The transmittance and reflectance are plotted as a function of the angle of incidence in Fig.
2.22. Note that, for parallel polarization, the reflectance vanishes and the transmittance is
€ angle.
unity at the Brewster

23
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

Figure 2.21: Calculation of the reflectance and transmittance coefficients.

(a)

(b)

Figure 2.22: Reflectance and transmittance for perpendicular (a) and parallel (b)
polarization directions, in the case of air/glass interface.

24
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

2.3.3 Total Internal Reflection and Evanescent Waves


It was shown that, for internal reflection with angles of incidence larger than the critical
angle, no real solutions exist for the transmission angles. To understand the physical nature of
the “transmitted” wave in this case we should have a closer look at the characteristics of the
electromagnetic field.

Figure 2.23: Propagation close to total internal reflection conditions at a planar interface.

Consider a plane wave impinging at an angle θi on a planar interface between a medium of


refractive index ni and a medium of index nt < n i (Fig. 2.23). The interface lies in the x-z
plane. The transmitted electric field is

E t = E0t e i( k t ⋅ r−ωt )

with

€ n
ktx =kt i sinθ i
n
t
k t ⋅ r = ktx x + k ty y = k t sin θ t ⋅ x + kt cos θ t ⋅ y

and


ni2
cosθ t = ± 1 − sin θ t = ± 1 − 2 sin 2 θ i
2

nt

where Snell’s law was invoked.


ni
For total internal reflection, we have sin θ t = sin θ i > 1, and thus
nt

n i2 2
cosθ t = ±i 2 sin θ i −1
€ nt

and hence

25
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

n i2 2
kty = ±ik t
sin θ i −1 ≡ ±iκ y
n t2
where we have introduced the real parameter

ni2 2
€ κ y = kt sin θ i −1
nt2

The transmitted electric field is therefore written

€ n
ikt i sin θ i x
i( ktx x ±iκ y y−ωt ) κ y y
E t = E0t e = E0t e e n t e−iωt

The physical solution corresponds only to the exponentially decaying field (along y): the
electric field is evanescent along the y direction (see Fig. 2.24). As a result, the field does not
carry any energy € flux (Poynting vector) perpendicular to the interface in the case of total
internal reflection. In fact, the field decays to 1/e of its value at the interface within a
(normal) distance
−1
1  n i2 sin 2 θ i 
Δy = = kt −1
κ y  n t2 

Figure 2.24: Evanescent wave obtained under the conditions of total internal reflection.

It is interesting to point out that, if a medium of refractive index at least as large as ni


is placed with its boundary parallel to the original interface, the transverse wave vector
becomes real again in that medium, and normal propagation in that medium resumes (see Fig.
2.25). This situation is referred to as frustrated total internal reflection (FTIR). It is
reminiscent of quantum mechanical tunneling of particles across finite-height potential
barriers. Such FTIR is useful for constructing, e.g., beam splitters, in which the reflection and
transmission coefficients can be tailored by tuning the low-index gap (see Fig. 2.26).

26
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

Figure 2.25: Frustrated total internal reflection (optical tunnelling).

Figure 2.26: Beam splitter based on frustrated total internal reflection (FTIR).

2.4 Geometrical and Paraxial Optics


Huygens’s and Fermat’s principles provide a strategy for evaluating the propagation of a
wave advancing with an arbitrary phase front, not necessarily propagating along a unique
axis. In many practical cases, however, the light rays propagate almost parallel to a given
axis, called the optical axis. A special case is that of geometrical optics, where the wave is
approximated by a single light ray, defining a unique direction of propagation. Another
special case, referred to as paraxial propagation, is encountered in many optical instruments
and devices in which the light wave propagates with its phase fronts approximately
perpendicular to the optical axis. We develop here the link between the wave picture of light
propagation and these two important limiting cases.

2.4.1 Relation between Wave and Ray Optics


For a monochromatic wave of vacuum wavelength λ0 = 2πc /ω propagating in a medium of
refractive index n(r ) (assumed to be slowly varying on the wavelength scale), the electric
field can be written in general as

E(r,t) = E(r )e iωt
with

E (r) = E0 (r)exp(−ik0 S(r))

27

E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

and k0 = 2π / λ0 = ω /c .
The function S(r ) is called the eikonal, and plays in optics the role of a potential in
classical mechanics. The eikonal is proportional to the phase, and hence surfaces of constant
value of S correspond to the wave fronts, whereas its gradient gives the directions of the light

rays (see Fig. 2.27):

S (r ) wave fronts
∇S ( r ) ray direction

Figure 2.27: The eikonal S and its relation with wavefronts and rays.

Notice that, for the special case of a plane wave, we have k0 S (r ) = k ⋅ r , and then

E (r) = E0 (r)exp (−ik ⋅ r).



We shall now derive the law of propagation for the light rays represented by ∇S(r) .
To this end, we apply the Helmholtz equation (Optics I ch. 1.3.2) to the electric field
expressed in terms of the eikonal:

[∇ 2
+ k 2 (r)]E (r) = 0

Note that this is an equation for the vector quantity E (r) , and hence represents three similar
€ E j (r); j = x, y,z
{ }
equations for the three components . For clarity, we consider each
component separately in the derivation that follows.
To calculate ∇ 2 E , we consider €the component ∇ 2 E j and write the Laplacian
explicitly: €

€ €
28
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

∇ 2E j ≡ [ ∇ ⋅ ∇ ]E j (r)
 ∂ ∂ ∂  ∂ ∂ ∂ 
=  ê x + ê y + ê z  ⋅  ê x + ê y + ê z E j (r)
 ∂ x ∂y ∂z   ∂ x ∂y ∂ z 

The first operation with the gradient operator on the field component yields (using
E (r) = E0 (r)exp(−ik0 S(r)) ):

 ∂ ∂ ∂
 ê x + ê y + ê z E0j (r)e−ik0S(r) 
€  ∂x ∂y ∂z 
∂E0j −ik S(r) ∂ S −ik0S(r) 
= ê x  e 0 − ik0E0j e 
 ∂x ∂x 

∂E0j −ik S(r) ∂ S −ik0S(r) 


+ê y  e 0 − ik0E0j e 
 ∂y ∂y 

∂E0j −ik S(r) ∂ S −ik0S(r) 


+ê z  e 0 − ik0E0j e 
 ∂z ∂z 

= e−ik0S(r)  ∇E0j (r) − ik0E0j (r)∇S(r)

The second operation with the gradient operator on this expression yields:

∇ 2E j = ∇ ⋅ e−ik0S(r)  ∇E0j (r) − ik0E0j ∇S(r)


{ }
 ∂ ∂ ∂
=  ê x + ê y + ê z  ⋅ e−ik0S(r)  ∇E0j (r) − ik0E0j ∇S(r)
{ }
 ∂x ∂y ∂z 
 ∂S ∂S ∂S 
= −ik0 e−ik0S(r)ê x − ik0 e−ik0S(r)ê y − ik0 e−ik0S(r)ê z  ⋅  ∇E0j (r) − ik0E0j (r)∇S(r)
 ∂x ∂y ∂z 

+e−ik0S(r) ∇ ⋅  ∇E0j (r) − ik0E0j (r)∇S(r)

= −ik0 e−ik0S(r) ∇S(r)⋅  ∇E0j (r) − ik0E0j (r)∇S(r) + e−ik0S(r) ∇ ⋅  ∇E0j (r) − ik0E0j ∇S(r)

The last term in the preceding expression can be calculated by writing the gradient operator
explicitly:

29
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

∇ ⋅  ∇E0j (r) − ik0E0j (r)∇S(r)


 ∂ ∂ ∂   ∂ ∂ ∂  ∂ ∂ ∂ 
=  ê x + ê y + ê z  ⋅  ê x + ê y + ê z E0j (r) − ik0E0j (r)  ê x + ê y + ê z  S(r)
 ∂x ∂y ∂ z   ∂ x ∂y ∂z   ∂x ∂y ∂z  
 ∂2 ∂2 ∂2   ∂E0j ∂E0j ∂E0j   ∂ S ∂S ∂S 
=  2 + 2 + 2 E0j − ik0  ê x + ê y + ê z  ⋅  ê x + ê y + ê z 
∂ x ∂ y ∂ z   ∂x ∂y ∂z   ∂ x ∂y ∂z 
 ∂ 2S ∂ 2S ∂ 2S 
−ik0E0j  2 + 2 + 2 
∂ x ∂ y ∂ z 
= ∇ 2E0j (r) − ik0 ∇E0j (r)⋅ ∇S(r) − ik0E0j (r)∇ 2 S(r)

Using this result in the previous expression obtained for ∇ 2 E j , we have:

∇ 2E j (r) = −ik0 e−ik0S(r) ∇S(r)⋅  ∇E0j (r) − ik0E0j (r)∇S(r) + e−ik0S(r) ∇ ⋅  ∇E0j (r) − ik0E0j (r)∇S(r)

= −ik0 e−ik0S(r) ∇S(r)⋅  ∇E0j (r) − ik0E0j (r)∇S(r)€

+e−ik0S(r) ∇ 2E0j (r) − ik0 ∇E0j (r)⋅ ∇S − ik0E0j (r)∇ 2 S(r)

= e−ik0S(r) ∇ 2E0j (r) − ik0E0j (r)∇ 2 S(r) − k02E0j (r) ∇S(r) − 2ik0 ∇E0j (r)⋅ ∇S(r)
2

2
Substituting into the Helmholtz equation, and using k 2 (r ) = k 0 n 2 (r ) , we obtain

[∇ 2
+ k 2 (r)]E0j (r)
2€
[
= e−ik0 S(r ) ∇ 2E0j (r) − ik0E0j (r)∇ 2 S(r) − k 0 E ]
0j (r) ∇S(r) − 2ik 0∇E0j (r) ⋅ ∇S(r) + k 0 n (r )E0j (r ) = 0
2 2 2

or:

∇ 2E0j (r) + k 0 n 2 (r )E0j (r) − k 0 E0j (r) ∇S(r) + i −2k 0∇E0j (r) ⋅ ∇S(r) − k 0E0j (r)∇ 2 S(r) = 0
2
2 2
{ }
The real part of this equation gives:

[ ]
∇ 2E0j (r) + k 0 n 2 (r ) − ∇S(r) E0j (r) = 0
2 2

or
€  λ 2
2
∇S E0j = n 2E0j +  0  ∇ 2E0j
 2π 

To obtain the case of ray optics, we take the limit of vanishing wavelength:
€  λ0  2 2
2
∇S E0j = n E0j +   ∇ E0j
2
→ n 2E0j
 2π  λ0 →0

30 €
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

This finally yields the wave equation in the geometrical optics limit, known as the eikonal
equation:

2
∇S (r ) ≅ n 2 (r )
or

∇S (r ) ≅ n (r )

Thus, the norm of the gradient of the eikonal at each point in space is approximately equal to
the refractive index at that local. For a given distribution of the refractive index n(r ) , solving

the eikonal equation yields the wave fronts S(r ) . Then, calculating back the gradient of the
eikonal ∇S (r ) finally gives the trajectory of the optical rays.
Note that the variation in the eikonal along a certain trajectory is related to the optical
path as defined in 2.2.2. To see this, we express the variation in S across a differential
segment ds along the ray trajectory (see Fig. 2.2.8) as

dS = ∇S (r ) ds ≅ n (r ) ds

Integrating between two points A and B along the trajectory yields


€ B rB

∇SAB = S(rB ) − S(rA ) = ∫ ∇S ds = ∫ n(r )ds


A rA

Thus, the variation in the eikonal along the ray trajectory is equal to the total optical path.
Recalling Fermat’s principle, it follows that ∇SAB /c , and hence ∇SAB , is stationary about the
actual trajectory.€

€ €

Figure 2.28: Integration path along a trajectory of a light ray for calculating the total
variation in the eikonal S.

2.4.2 Paraxial Approximation


The paraxial approximation of the wave equation and its solutions represent a common case
often encountered in optical systems, namely, that of optical beams having wavefronts that
are nearly perpendicular to a fixed direction in space. In terms of a ray picture, this
corresponds to optical rays that propagate nearly parallel to the propagation (optical) axis.

31
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

The case of paraxial propagation can be formulated starting from the scalar spherical
wave U (r ) = U 0 e − ikr / r , which is a solution of the Helmholtz equation

(∇ 2 + k 2 )U (r ) = 0

with the wavenumber k = 2π / λ . The coordinate r represents the distance from the point
source of the spherical wave to the point where its amplitude is U(r) (see Fig. 2.29(b)). In
the paraxial approximation, the rays stay close to the principal axis of propagation, which is
defined here as the z-direction. We can therefore write:
€ 1
1  x2 + y2  2
r = (x + y + z
2 2 2
) 2
= z1+ 
 z2 
 x2 + y 2  x2 + y 2
≈ z1 +  = z +
 2z2  2z

and hence the spherical wave solution can be approximated by

x2 +y2
e− ikz − ik
2z
U( r) ≈ U 0 e
x2 + y2
z+
2z

The denominator in this expression is

1 1  x2 + y 2  1 x2 + y 2 1
≈ 1 −  = − ≈
x 2 + y 2 z  2z2  z 2 z3 z
z+
2z

and we finally obtain:

x2 +y2
U − ik
U(r) ≈ 0 e− ikze 2z
≡ A(r)e−ikz
z

The amplitude of the scalar field is thus given by

x 2 +y 2
U −ik
A(r) ≡ 0 e 2z

and is referred to as a paraboloidal wave, since its constant phase surfaces consist of
paraboloids of revolution (about the z axis). Note that the main z-dependence of the solution
−ikz
U is contained in the phase factor e whereas the envelope function A(r) exhibits relatively
slow variations of the amplitude along z. The phase fronts of a paraboloidal wave are
sketched in Fig. 2.29(c). They essentially consist of the segments of a spherical wave cut by a
cone that subtends a small angle about the optical axis, where the paraxial approximation
holds. €

32
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

The form of the Helmholtz equation in the paraxial approximation is obtained by


rewriting the Laplacian operator as:

∂2
∇ 2 = ∇ 2⊥ +
∂z 2

Using

∂2 ∂2 ∂2 A − ikz ∂A −ikz
2 U (r) ≈ 2 A(r)e −ikz
= 2 e − k 2 A(r)e −ikz − 2ik e
∂z ∂z ∂z ∂z

and the fact that the z-variation of A(r) is relatively slow over a distance of a wavelength,
i.e.,

∂2 A ∂A
€ 2
<< k << k 2 A
∂z ∂z

we substitute into the Helmholtz equation:

∂A −ikz
0 = ∇ 2U + k 2U ≅ e−ikz∇ 2⊥ A − 2ik e
∂z

and finally obtain the paraxial version of the scalar wave equation:
€ ∂A(r)
∇ 2⊥ A(r) − 2ik ≈0
∂z

Paraboloidal waves are solutions of this equation, as can be verified by direct substitution.

33
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

Figure 2.29: Schematic illustration of the wavefronts of (a) plane wave; (b) spherical wave;
(c) paraboloidal wave.

2.4.3 Application: Gaussian Beams


The paraboloidal waves are but one solution for the paraxial wave equation. Another
important class of paraxial solutions comprises Gaussian beams. These solutions can be
derived by formally replacing the coordinate z by a complex number:

z
→ z + iz0 ≡ q ( z)

The envelope function then becomes

x2 + y 2
U0 −ik
A(r ) 
→ e 2 ( z+iz0 )
z + iz0
We now rewrite

1 z − iz 1 1
= 2 02 = 2 −i 2
q ( z) z + z0 ( z + z0 ) z ( z + z02 ) z0
2

34
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

and, by introducing the parameters

 z 2 + z02   z0  2 
 R( z) ≡ = z1 +   
 z   z  

 2 z 2 + z02 λ λz0  z 2    z 2 
2
W ( z) ≡ = 1 + 2  ≡ W0 1 +   
 z0 π π  z0    z0  

W ≡ λz0
 0 π

we obtain:

1 1 λ
= −i
q ( z) R( z) πW 2 ( z)
Using these newly defined parameters, the scalar wave solution, termed a Gaussian beam,
reads:

  A0

 ρ2
 U0  W0 −ρ 2 W 2 (z )
−ikz−ik
2R(z)
+iζ (z )
 U(r) =   e ⋅e
 z0  W (z)

ζ (z) = tan−1 (z z0 )

where ρ 2 ≡ x 2 + y 2 .
The phase fronts and the transverse extension of a Gaussian beam propagating along

the z-axis are represented schematically in Fig. 2.30.

Figure 2.30: Schematic representation of a Gaussian beam and its phase fronts.

35
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

Figure 2.31: Illustrations of the parameters of a Gaussian beam.

By direct inspection of the mathematical expressions defining the Gaussian beam, one
can derive its important features. Such beam exhibits Gaussian intensity distributions within
the x-y plane at each point z along its propagation axis (Fig. 2.31). It is further characterized
by several parameters:

(i) The beam width:


 z 2
W (z) = W 0 1+   z
→ ∞
→ ∞
 z0 

defines the contours in the x-y plane that entrap 86% of the beam intensity (Fig.
2.32). The minimal beam width, located at z=0, is called the beam waist:

W (z = 0) = W0

(ii) The radius of curvature of the phase fronts is given by R(z) (Fig. 2.33). It is
minimal at z = ±z0 :

R( z = z0 ) = 2z0

It tends to infinity at z = 0 (planar phase fronts) and approaches the value for a
spherical wave centered at z=0 at large distances from the waist:

R(z) → z
€ z→∞

(iii) The Rayleigh range z0 represents the zone where the beam remains collimated
(i.e., of width near 2W0):
2
2z0 = 2π W0 / λ

36
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

(see Fig. 2.34).

Figure 2.32: Variation of the width of a Gaussian beam along its propagation axis.

Figure 2.33: Variation of the radius of curvature of the phase fronts of a Gaussian beam
along its propagation axis.

Figure 2.34: Width of a Gaussian beam along its propagation axis, indicating the Rayleigh
range z0.

It is interesting to note that the divergence (half) angle of a Gaussian beam is given by
(see Fig. 2.32)

W ( z) 1 1 W
θ0 ≅ = W0 2 + 2 
z
→ 0 = const

z z z0 → ∞
z0

37
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

or, using the expression for the Rayleigh range,

λ
θ0 =
W0 π

Thus, the product of the minimum cross section of the beam (at z = 0):

A0 = πW 02

and the solid divergence angle:

Ω 0 = πθ 02

is constant:

A0 Ω0 = (πW 02 ) ⋅ (πθ 20 ) = λ2

In this sense, a Gaussian beam represents the case of a minimum uncertainty in a


beam (waist) size and direction. On the other hand, two limiting cases of a Gaussian beam,
namely a spherical wave and a plane wave, represent the two extreme cases of this
uncertainty. A spherical wave represents the case of a maximum uncertainty in propagation
direction (isotropic) and zero uncertainty of its waist size (point source). On the other hand, a
plane wave corresponds to zero uncertainty in propagation direction (along the wave vector)
and maximum uncertainty in “waist” size.
Finally, it is worth noting that Gaussian beams represent a class of “real” waves, as
opposed to plane waves, which are useful more as a mathematical tool. This is because the
extent of a Gaussian beam is finite and thus such beam carries a finite amount of energy. In
fact, Gaussian beams are a special case of Gauss-Hermite modes, the optical fields generated
by certain optical cavities often used in laser instruments. This will be further discussed in
Optics II.

Gaussian Beam Focusing:

Since a Gaussian beam can be taken as a good approximation of the “real” beam generated
for example by a laser, it can be interesting to see how its properties can be manipulated by
inserting a thin lens into its path. First of all, we need to find out how a plane wave
propagates through a lens composed of a dielectric material having refractive index n and
maximum thickness d0 (see Fig. 2.35). As it turns out, the effects of the presence of such a
lens can be modeled through the introduction of a spatially dependent phase difference,
caused by the continuously varying lens thickness. In order to define this phase difference,
we must then write an expression for the length δd, as defined in Fig. 2.35:

  2 2  2
1
x + y
δd( x, y ) 2
= R− R − x + y ( 2
= R 1− 1−
2
)  
  R2  
 
 2 2 2 2
x +y  x +y
≈ R 1 −1+ =
 2R 2  2R

38

E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

R
From this expression, and from the fact that f = (where f is the focal length of the
n −1
lens), we can estimate the distance d traveled in the dielectric by the wavelet impinging on
the lens in the (x,y) point,
x2 + y2 x2 + y2
⇒ d ( x, y ) = d0 − δ€d = d0 − = d0 − .
2R 2 f ( n −1)

The different phase picked up by each wavelet in passing through the lens results in a
transmission coefficient, t, which is proportional to

k0 2 2
−i( n−1)k0 d ( x,y) −i( n−1)k0 d0
+i
2f
(x +y )
t ( x, y) ∝ e =e e .

Figure 2.35: Propagation of a plane wave through a thin lens: definition of relevant
parameters. R, n, and d0 are, respectively, the radius of curvature, the refractive index, and
the maximum thickness of the lens.

The same expression of t holds for a double structure (Fig. 2.36), since in this case

 d
x2 + y2 x2 + y2
d ( x, y ) ≅ d1 + d2 − −
2R1 2R2
1 1 1 1 1
+ = ( n −1) +  ≡
f1 f 2  R1 R2  f
€ x2 + y2
⇒ d ( x, y ) = d −
2 f ( n −1)

and

39
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

ik0
x 2 +y 2 1 1 1
t ( x, y ) ∝ e 2f ; = ( n −1)  − 
f  R1 R2 

Figure 2.36: Double-lens structure.

Before we proceed further, we would like to remind the reader that a Gaussian beam can be
written as

A0
  ρ2
 U 0  W 0 − ρ 2 W 2 (z) −ikz−ik 2R(z) +iζ (z)
U(r) =   e ⋅e .
 z0  W (z)

In passing through a lens (Fig. 2.37), the phase of such a beam changes in the following way:

ρ2 ρ2 ρ2
−ikz − ik + iζ (z) → −ikz − ik + iζ (z) + ik
2R(z) 2R(z) 2f
kρ 2  1 1
= −ikz − i  −  + iζ (z).
2  R(z) f 

Therefore, the presence of the lens introduces a variation in the radius of curvature R of the
beam, which becomes

−1
 1 1
R' ( z) =  −  .
 R(z) f 

Figure 2.37: Gaussian beam focusing.

40
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

 1 1
If R>f,  −  < 0 , so that R' ( z) < 0 . This implies that our Gaussian beam will converge
 R(z) f 
after going through the lens, acquiring its minimum width 2W 0 ' at a distance z’ from the lens.
We can calculate the new waist size and distance from the lens in the following manner.
Given that for the incoming
€ beam W and R can be written as

 1€
  2 2
 z  λz0

W ( z) = W 0 1+  z   ; W 02 =
  0  π

   z 2
 R( z) = z 1+  0  
   z  

we can reconstruct z in the following way:



  2    2 
z λ z
W = W 1+    =
2
0
2
z0 1+   
 z   π   z0  
 0 
 2
 π 2 z2 π2 4 ( z02 + z 2 )
 λ W = z0 + z ⇒ λ 2 W = z02
⇒  0

 R z2 R 
 = 1+ 02 ⇒ z02 =  −1 ⋅ z 2
 z z z 
2
 R  2 2 
 −1 z + z 
π2 4  z   R2
⇒ W = =
λ2 R  2 R 
 −1 z  −1
z  z 

R R2 λ 2 R2
⇒ −1 = =
z π2 4 π 2W 4
W
λ2
z 1 R
⇒ = ⇒ z= 2
R R2λ 2  λR 
1+ 2 4 1+  
π W  πW 2 

while for W0 we have

41
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

2 2 R2 R2
z 0 = zR − z = 2
− 2
 λR    λ R 2 
1+   1+  
 πW 2    π W 2  
2 2
2  λR 
2 2  λR 
2
R +R   −R R  
 πW 2   πW 2  π2 4
= 2
= 2
= W0
  λ R 2    λ R 2  λ2
1+   1+  
  π W 2     π W 2  
4
λ4
4
R4λ 4  πW 2 
R 4 4  
π W π 4W 4  λ R  W4
W04 = 2
= 2
= 2
  λ R 2    π W 2 2    π W 2 2 
1+   1+   1+  
  π W 2     λ R     λ R  

W
⇒ W0 = 1/2
.
  π W 2 2 
1+  
  λ R  

According to what we saw previously, on the right-hand side of the lens we have that
 −1
 R → R' =  1 − 1 
 R f  ,
 W → W ' = W

and for the emerging Gaussian beam we can write


€ W
W '0 = 1
(waist)   2 2  2
1+  πW  
  λR'  
R'
−z' =
(focal point)  λR'  2
1+  
€  πW 2 

Example: Apodized plane wave

The formalism developed above €can be used to describe the effects of a thin lens on a
Gaussian apodized plane wave. Like a homogeneous plane wave, this particular kind of wave
has a well-defined propagation direction (along its wave vector), but it is characterized by a
Gaussian intensity profile across its wave front (see Fig. 2.38). For this particular case, we
have that

 R' = − f πW 2 πW 0 2 z
 ⇒ = − = − 0
W = W 0 λR' λf f

42

E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

and also that z = 0; 2z0 >> f . As a consequence, for the beam waist after the lens ( W 0 ' ) we
can write

W0 f f
€ W0 ' = 1
≈ W0 = W0 €
  z 2  2 z0 π W02
1+  0   λ
  f  
λf  W0 ' 
= = θ0 f  ≅ θ0 
π W0  f 

while for the focal point z’ we have (since R’=-f)

f   f 2 
z' = ≈ f 1+    ≈ f .
 f 2   z0  
1+  
 z0 

The lens diameter, Dlens, should at least be equal to 2W0 to achieve maximum focusing. If we
assume Dlens = 2W 0 , the waist of the focused beam is

λf 2λ
W0' = 2 = ⋅ F# ,
πD π

f
where F# = is the F-number of the lens.
d

Figure 2.38: Focusing of an apodized plane wave (note the Gaussian intensity profile in the
x-y plane).

43
E. Kapon/Optics I (2016/2017)/Chapter 2: Propagation of Light

2.5 Summary

ε (ω ) Ne2 1
-Refractive index dispersion = n˜ 2 (ω ) = 1 +
ε0 ε 0 m ω 0 − ω 2 + iσω
2

-Snell’s law ni sinθ i = nt sinθ t

-Total internal reflection θc = sin−1 (n t /n i )

-Brewster’s angle tan θ B = n t /n i

 sinθ t cos θi − cosθ t sin θ i sin(θ i − θ t )


r⊥ = =−
 sin θ t cosθ i + cosθ t sinθ i sin(θ t + θ i)
-Fresnel equations (s-polarization) € 
2cosθ i sin θt 2cosθ i sin θt
t ⊥ = =
 sin θi cos θi + cos θt sinθ t sin(θ i + θ t )

 cosθ i sinθ i − cos θt sinθ t tan(θ i − θ t )


r = =
 // cos θt sinθ t + sin θ i cosθ i tan(θ i + θ t )
-Fresnel equations (p-polarization) 
2cosθ i sinθ t 2sinθ t cos θi
t // = =
 cosθ t sinθ t + cos θi sin θ i sin(θ i + θt ) sin(θ i − θ t )

E(r ) = E0 (r)e
−ik0 S( r )
-Eikonal

2  λ0  2 2
-Eikonal equation ∇S E0 = n E +   ∇ E
2
→ n 2E
 2π  λ0 →0

x2 +y2
U − ik
-Parabolidal wave U(r) ≈ 0 e− ikze 2z
≡ A(r)e−ikz
€ z

∂A
-Paraxial approximation ∇ 2⊥ A − 2ik ≈0
∂z

  A0

 ρ2
 U0  W0 −ρ 2 W 2 (z )
−ikz−ik
2R(z)
+iζ (z )
-Gaussian beam U(r) =  z  W (z) e ⋅e
  0
ζ (z) = tan−1 (z z0 )

-Divergence of a Gaussian beam A0Ω0 = λ2


44

You might also like