Fotonica Book Chapter 6
Fotonica Book Chapter 6
Fotonica Book Chapter 6
Electromagnetic Optics
Contents
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6–1
6.2 Maxwell’s electromagnetic wave equations . . . . . . . . . . . . . . . . . . . . . . 6–2
6.3 Dielectric media . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6–3
6.4 Elementary electromagnetic waves . . . . . . . . . . . . . . . . . . . . . . . . . . . 6–5
6.5 Polarization of electromagnetic waves . . . . . . . . . . . . . . . . . . . . . . . . . 6–8
6.6 Reflection and refraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6–11
6.7 Absorption and dispersion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6–15
6.8 Layered structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6–16
6.9 Scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6–25
6.1 Introduction
Light is an electromagnetic wave phenomenon that is described by the same theoretical principles
used for all electromagnetic radiation. Light or optical radiation (or optical frequencies) are all
frequencies between infrared, visible and ultraviolet light, so all wavelengths (roughly) between
10nm and 1mm. Propagation of electromagnetic radiation is expressed by two coupled symmet-
rical partial differential equations, coupling the electric field vector with the magnetic field vector.
These equations were originally formulated by James Clark Maxwell in 1864. Maxwell’s theory
was not only a breakthrough in physics because it was the first example of unification (magnetism
and electricity, at first sight separate phenomena, appeared to be fundamentally linked), but also
because it led Einstein directly to his theory of relativity. From Maxwell’s laws it follows that the
speed of light is always 299792458 m/s. However, according to classical physics velocities can be
added, so a light ray emitted by a fast object would have a speed larger than 299792458 m/s. This
paradox made Einstein think, resulting in his famous theory of relativity.
The scalar wave optics theory discussed in chapter 4 is an approximation of Maxwell’s equations,
because light is described by one single scalar wave equation. This single scalar equation is suf-
ficient for the paraxial approximation with certain conditions (explained later). By performing
6–1
another approximation, the short wavelength limit, we already arrived at geometrical optics, see
chapter 3.
In this chapter we present a short overview of the important aspects of electromagnetic theory for
optics. We start from Maxwell’s equations and discuss some elementary waves. Then we describe
properties of dielectric media. These two sections form the postulates of electromagnetic optics: a
set of rules for the next sections. Furthermore we discuss polarization, absorption and dispersion,
and the laws of reflection and refraction. To conclude a few layered structures are examined.
The electric and magnetic field vectors E(r, t) (unit: V /m) and H(r, t) (unit: A/m) in a medium
without free charges or currents, satisfy the following coupled partial differential equations which
are function of space r and time t: Maxwell’s equations.
∂D
∇×H =
∂t
∂B
∇×E = −
∂t
∇·D = 0
∇·B = 0 (6.1)
The vector fields D(r, t) (unit: C/m2 ) and B(r, t) (unit: W b/m2 ) are the electric flux density (also
called electric displacement vector or electrical induction) and the magnetic flux density (also
called magnetic induction), respectively. The relation between D and E depends on the electrical
properties of the medium. Analogously, the relation between B and H depends on the magnetic
properties. They form the constitutive relations:
D = 0 E + P
B = µ0 H + µ0 M (6.2)
The constants µo = 4π10−7 H/m and 0 = c21µ0 F/m are the permeability and the permittivity of the
vacuum, respectively. P (unit: C/m2 ) is the polarization density, and M (unit: A/m) is the mag-
netization density. In a dielectric medium the polarization density P is equal to the macroscopic
sum of the electric dipole moments induced by the electric field. An analogous definition can be
given for M. Further on we will see that the fields P and M are related to E and H, respectively,
by relations dependent on the electrical and magnetic properties of the material.
In free space (= non-electrical and non-magnetic) we have: P = M = 0, so D = 0 E and B = µ0 H.
Notice that in this case the Maxwell equations are reduced and decoupled to the scalar wave
equation for all three vector components, because the permittivity or refractive index is constant.
P=E×H (6.3)
6–2
known as the Poynting vector. The power follows the direction of this vector, that is perpendicular
to both E and H. The optical intensity1 I, which is the power per surface area perpendicular to P,
is equal to the magnitude of the Poynting vector, averaged over a certain time, see section 4.1.2.
The energy density (unit: J/m3 ) associated with an electromagnetic wave is given by
U = (E · D + H · B)/2 (6.4)
The first and second term represent the energy carried by the electric field and the magnetic field
respectively.
It is convenient to view the medium equation (eq. 6.2) between E and P as a system where the
medium responds to an applied electric field E (input) and creates a polarization density P as
output or response. We give a few definitions relating to dielectric media. A dielectric is:
• Linear: If P(r, t) is linearly related to E(r, t). Then the superposition principle applies.
• Isotropic: If the relation between P(r, t) and E(r, t) is independent of the direction of E(r, t),
so the medium looks the same from all directions. Then, the vectors E(r, t) and P(r, t) have
to be parallel.
• Spatially non-dispersive: If the relation between P(r, t) and E(r, t) is local; if P(r, t) at
location r is only influenced by E(r, t) at the same position r. In this chapter we assume that
all media are spatially non-dispersive.
In this chapter we will use non-magnetic materials (M = 0) without free electrical charges or
currents. In addition, if the medium is linear, non-dispersive, homogeneous and isotropic we get:
P = 0 χE. (6.5)
Here the scalar constant χ is the electric susceptibility. It follows that P and E are parallel at each
position and time, just like D and E:
D = E (6.6)
1
The use of the term ‘Intensity’ is a mess in optics. The term is used on the one hand for optical power density
(W/m2 ), but also for electric field energy density (J/m3 ). To make matters worse the term is widely used in radiometry
and photometry to denote radiant intensity (W/str) or luminous intensity (Candela). In all cases it has ‘something’ to
do with power or energy.
6–3
with
= 0 (1 + χ) (6.7)
The scalar constant is the electrical permittivity of the medium. With the previous conditions the
Maxwell equations reduce to:
∂E
∇×H =
∂t
∂H
∇ × E = −µ0
∂t
∇·E = 0
∇·H = 0 (6.8)
Note that the equations are reduced and decoupled to the scalar wave equation for each of the
three components of E and H:
1 ∂2u 1
∇2 u − =0 with v 2 = (6.9)
v 2 ∂t2 µ0
The components of the electric and magnetic field propagate in the medium with velocity v, ac-
cording to:
c
v = (6.10)
n
r
p
n = = (1 + χ) (6.11)
0
with c the speed of light in free space. The constant n is equal to the ratio of the speed of light in
free space to the speed in the medium. It is called the refractive index of the material.
Boundary conditions at an interface The boundary conditions at an interface between two lin-
ear, isotropic, homogeneous and non-magnetic media with dielectric constants 1 and 2 , are im-
portant. We get:
n × (E1 − E2 ) = 0 (6.12)
n × (H1 − H2 ) = 0 (6.13)
n · (1 E1 − 2 E2 ) = 0 (6.14)
n · (B1 − B2 ) = 0 (6.15)
The tangential components of the electric and magnetic field, and the normal component of the
magnetic field, are continuous. The normal component of the electric field makes a discontinuous
jump.
In a non-homogeneous medium the electrical susceptibility, the dielectric constant and thus re-
fractive index are a function of the position r. An example of a non-homogeneous medium is a
6–4
graded index medium. One can prove (by using ∇× on the Maxwell equations) that the scalar
wave equation of eq. (6.9) obtains an extra term:
1 ∂2E
2 1
∇ E− 2 +∇ ∇ (r) .E = 0 (6.16)
c (r) ∂t2 (r)
Notice that the location dependent refractive index results in a location dependent speed of the
wave in the medium.
For locally homogeneous media, so (r) varies slowly in space, the third term on the left side can
be neglected.
In dispersive media E will create P by inducing oscillations of bound electrons in atoms of the
medium, so they can collectively and with a certain retardation build up a polarization density.
Because we assume a linear medium, an arbitrary electric field will induce a polarization density
P (t)composed of the superposition of all E(t0 ) with t0 < t, or:
+∞
Z
χ t − t0 E t0 dt0
P (t) = 0 (6.17)
−∞
which is a convolution integral, with 0 χ(t) the polarization density response to an impulse of
electric field.
A monochromatic plane wave is a wave where all components of the electric and magnetic field
are harmonic functions in time with the same frequency. To simplify notations these components
are presented with their complex amplitudes, as in section 4.2
E(r, t) = Re E(r)ejωt
(6.18)
jωt
H(r, t) = Re H(r)e (6.19)
Here E(r) and H(r) are the complex amplitudes of the electric and magnetic field. In the same
way the complex amplitudes of P(r, t), D(r, t) and B(r, t) are denoted as: P(r), D(r) and B(r).
Maxwell’s equations (for linear, non-dispersive, homogeneous and isotropic media) for monochro-
matic waves are derived by substitution of the complex amplitudes in (6.8). If we also perform the
6–5
substitution: ∂/∂t = jω we obtain:
∇ × H = jωE (6.20)
∇ × E = −jωµ0 H (6.21)
∇·E = 0 (6.22)
∇·H = 0 (6.23)
P (r) = 0 χ(r, ω)E(r) (6.24)
(r, ω) = 0 (1 + χ(r, ω)) (6.25)
p (r, ω)
n(r, ω) = ( ) (6.26)
0
(6.27)
We already know that the electromagnetic power flux is equal to the time averaged Poynting
vector. With complex amplitudes we get:
1 1
P = Re Eejωt × Re Hejωt = (Eejωt + E∗ e−jωt ) × (Hejωt + H∗ e−jωt )
2 2
1
= (E × H∗ + E∗ × H + E × He2jωt + E∗ × H∗ e−2jωt ) (6.28)
4
By averaging over time the exponential terms will disappear and we obtain:
1 1
hPi = (E × H∗ + E∗ × H) = (S + S∗ ) = Re {S} (6.29)
4 2
with
1
S = (E × H∗ ) (6.30)
2
The vector S is called the complex Poynting vector. The optical intensity is equal to the magnitude
of the vector Re {S}.
We consider a monochromatic plane wave in a medium (without sources) that is linear, non-
dispersive, homogeneous and isotropic. For the electric and magnetic components with wave
vector k we have the complex amplitudes:
6–6
Figure 6.1: TEM plane wave. The vectors E, H and k are perpendicular. The wavefronts are normal to k.
previous amplitudes in the first two Maxwell equations (6.20) and (6.21) in the frequency domain,
we get:
k × H0 = −ωE0 (6.33)
k × E0 = ωµ0 H0 (6.34)
An example of an electromagnetic spherical wave is the field radiated by an electrical dipole. Such
a spherical wave can be constructed by use of an auxiliary field A:
1
U (r) = e−jkr (6.39)
r
where ex is the unit vector along the x-direction and also represents the direction of the dipole. We
know that U (r) satisfies the Helmholtz equation (see chapter 4), so A(r) is also a solution of the
6–7
Helmholtz equation, and it is called the electromagnetic vector potential. It can be proven that:
1
H = ∇×A (6.40)
µ0
1
E = ∇×H (6.41)
jω
The concept ‘polarization’ relates to the fact that the orientation of the electric field vector E(r, t)
of an electromagnetic wave changes in time if we look at the vector at a certain location in space.
The state of polarization is completely known if we know how the orientation of the electric field
vector changes in time.
The polarization of light has important consequences for the interaction of light with matter:
• The amount of reflected light at an interface depends on the polarization of the incident
wave.
• The refractive index of anisotropic materials depends on polarization. Waves with different
polarization propagate with different speeds, thus they experience different phase changes
so the polarization ellipse (see further) will transform.
Consider a monochromatic plane wave with frequency ν propagating in the z-direction with
speed c. The electric field is in the xy-plane and is described in general by:
n z
o
E(z, t) = Re Aej2πν(t− c ) (6.42)
Starting from the real representation of a monochromatic wave at a certain location in space we
can see that the most general movement of the electric field vector in time is an ellipse. We call
this an elliptical polarization state.
We write Ax and Ay with their magnitude and phase Ax = ax ejφx , Ay = ay ejφy . We substitute this
into eq. 6.42 and obtain:
E(z, t) = Ex ex + Ey ey (6.44)
6–8
Figure 6.2: Elliptically polarized light. (a) Rotation of the end point of the electrical field vector in the
xy-plane at a fixed location in space. (b) Trajectory in space at a fixed time t.
with
h z i
Ex = ax cos 2πν t − + φx
c
h z i
Ey = ay cos 2πν t − + φy (6.45)
c
The components Ex and Ey are periodic functions of (t − z/c) and oscillate with frequency ν.
These equations are the parameter equations of an ellipse. Indeed, by eliminating t we get:
Ex2 Ey2 Ex Ey
+ 2 − 2 cos φ = sin2 φ (6.46)
a2x ay ax ay
Here φ = φy − φx is the phase difference. At a fixed location z in space the end point of the
electric field vector will rotate periodically in the xy-plane describing an elliptical trajectory, see
figure 6.2a. At a fixed time t the location of the end point will follow a helical trajectory in space,
see figure 6.2b. However when we travel along with the field at the speed of light ( t − zc =
constant), we will always see the same field orientation.
The complete state of polarization is known if we know the plane of the ellipse, the direction and
magnitude of the main axes, the direction of revolution and the starting phase (thus the orientation
of the electric field at time t = 0).
If for elliptical polarization one of the components is dropped, e.g. ax = 0, then the light is linearly
polarized in the direction of the other component (e.g. y-direction). Light is also linearly polarized
if the phase difference φ = 0 or π, because then we obtain from eq. 6.46: Ey = ±(ay /ax )Ex . This is
the equation of a line with slope ±ay /ax . These cases at a fixed position z and at a fixed time t are
shown in figure 6.3.
6–9
Figure 6.3: Linearly polarized light. (a) Time evolution at a fixed point in space. (b) Space evolution at a
fixed time t.
If φ = ±π/2 and ax = ay = a0 , then we obtain from eq. 6.46: Ex2 + Ey2 = a20 , which represents a
circle. The elliptical cylinder of figure 6.2 will now be a circular cylinder and the wave is circularly
polarized. If φ = +π/2, the field at a fixed position z rotates clockwise, viewed from the direction
the wave is propagating to. This is called right-hand circular polarization. The case φ = −π/2
corresponds to a left-hand circularly polarized wave.
Unfortunately right- and left-handed polarization is not univocally defined in literature. In op-
tics and physics the definition is used that right-handed corresponds to a clockwise movement
when one looks into the bundle, while in the world of radiowaves and microwaves the reversed
definition is used.
It is clear that the E-vector of an elliptical wave can be considered as the superposition of 2 linearly
polarized waves. Because of linearity of Maxwell’s equations, this means that the analysis of an
optical system w.r.t. all possible polarizations can be limited to the behavior for 2 orthogonal linear
polarizations.
As was explained in the chapter on scalar waves, the superposition of two (or more) waves leads
to interference effects in the intensity of those waves. At optical frequencies this means that an
optical detector can ‘see’ fluctuations in the detected intensity: in certain locations in space there
may be destructive interference and no signal is picked up by the detector, whereas in others there
is constructive interference and therefore a strong signal is picked up. For most optical detectors
(and also for the human eye) the relevant intensity is the energy density of the electric field, as
given by (E · E)/2. If two fields E1 and E2 are present, the total energy density is given by
(E · E)/2 = (E1 + E2 ) · (E1 + E2 )/2 = |E1 |2 /2 + |E2 |2 /2 + E1 · E2 (6.47)
6–10
Figure 6.4: The problem of an electromagnetic wave (a) incident on an interface can be separated into a
TE-problem (b) and a TM-problem (c). Both are decoupled.
From this expression it is clear that interference fringes will only be detectable if the constituting
fields are not orthogonal. More in particular orthogonal polarizations will never interfere!
In this section we examine reflection and refraction of a monochromatic plane wave with arbitrary
polarization, incident on a plane interface between two dielectrics. We assume that these media
are linear, homogeneous, isotropic, non-dispersive and non-magnetic. Figure 6.4 and 6.5 present
an overview of the problem: we have two media with indices n and n0 , an incident wave, a re-
flected wave and a refracted wave. Already in chapter 4 we showed that the wave fronts of the
incident and the reflected wave agree at the interface only if θ = θ00 . Snell’s law was also obtained:
nsinθ = n0 sinθ0 .
Now we want to get the reflection and transmission coefficient for the reflected and refracted
wave. Therefore we demand that the fields satisfy the boundary conditions at the interface. Pre-
viously we observed that the tangential components of E and H, and the normal components of
D and B, have to be continuous at the boundary. Furthermore, we noted that the ratio of the am-
plitude of the magnetic field to the
p perpendicular electric field is equal to E/H = Z0 /n, with Z0
the free space impedance (Z0 = µ0 /0 ), and with n the refractive index of the medium in which
the wave is propagating.
When solving Maxwell’s equations at the interface, the problem reduces to a two-dimensional
one, because the fields at an interface are y-invariant, see figure 6.4. One can prove (substitution of
two-dimensional fields in Maxwell’s equations) that the general solution of the equations for two-
dimensional phenomena are separated into two partial problems: we get two decoupled sets of
differential equations. One gives the solution for the components: Ey (x, z), Hx (x, z) and Hz (x, z).
These are called TE or transversal electric solutions (sometimes also called s-polarization), because
the single component of the electric field is transversal (=perpendicular) to the plane of incidence
(being the plane containing the direction of incidence, and which is normal to the interface). The
other differential equation set determines: Hy (x, z), Ex (x, z) and Ez (x, z), which are analogously
called TM or transversal magnetic solutions (or sometimes p-polarization).
6–11
From the previous data it is possible to calculate the reflection and transmission coefficients for
both TE and TM polarizations (do this yourself!). The results are:
6.6.1 TE polarization
External reflection (n0 > n). The reflection coefficient rT E is always real and negative, which
corresponds to a phase shift φT E = π. The magnitude |rT E | for perpendicular incidence (θ = 0) is
n0 −n ◦
equal to n+n 0 . For θ = 90 , |rT E | = 1.
Internal reflection (n0 < n). For small θ the reflection coefficient rT E is real and positive. The
n−n0
magnitude |rT E | for perpendicular incidence (θ = 0) is n+n 0 . At a certain angle θ we get that
6–12
Figure 6.5: Magnitude and phase of the reflection coefficient in function of incidence angle for (a) exter-
nal reflection (n0 /n = 1.5) and TE polarization, (b) external reflection (n0 /n = 1.5) and TM polarization,
(c) internal reflection (n/n0 = 1.5) and TE polarization and (d) internal reflection (n/n0 = 1.5) and TM
polarization.
Exercise: Derive an expression for the decay constant of the tail in the TIR regime as a function
of angle of incidence.
6.6.2 TM polarization
External reflection (n0 > n). The reflection coefficient rT M is real. The magnitude |rT M | for per-
n0 −n
pendicular incidence (θ = 0) is equal to n+n 0 and decreases for increasing θ, until |rT M | = 0. This
6–13
Figure 6.6: Reflectance for TE and TM polarization at an interface between air and GaAs (n0 = 3.6).
For θ > θB rT M will change sign and its magnitude increases gradually until it reaches 1 at θ =
90◦ . The fact that a TM-wave is not reflected at the Brewster angle is used for the fabrication of
polarizers (devices that block a certain polarization and transmit another).
Internal reflection Analogous discussion.
The reflection and transmission coefficients r and t are ratios of complex field amplitudes. The
power reflection (or reflectance) R and power transmission (or transmittance) T is defined as the
ratio of optical flux densities (along a direction perpendicular to the surface) of reflected and trans-
mitted wave, relative to the incident wave. Because the incident and reflected wave propagate in
the same medium, and their angles with the interface are the same, we obtain:
R = |r|2 . (6.58)
n0 cos θ0 2
Power conservation dictates: T = 1 − R. Note that T = n cos θ |t| , which is not equal to |t|2 , as the
power propagates along a different angle.
An important case is that of perpendicular incidence on an interface. The reflectance, resp. trans-
mittance, is the same for TE and TM, both for internal and external reflection, and is equal to:
n − n0 2
R =
n + n0
4nn0
T =
(n + n0 )2
Example: the reflectance and transmittance at the interface between glass (n0 = 1.5) and air is
4% for perpendicular incidence. Figure 6.6 shows the reflectance for TE and TM between air and
GaAs (n0 = 3.6) in function of the incidence angle θ.
6–14
6.7 Absorption and dispersion
6.7.1 Absorption
Up until now we assumed that the dielectric media were completely transparent, there was no
absorption of light by the material. For example, glass is very transparent in the visible part of the
spectrum, but it strongly absorbs infrared and ultraviolet light. Dielectrics that absorb light are
often described by a complex susceptibility:
χ = χR + jχI (6.59)
The ratio between optical power after a certain propagation (Po ) and initial optical power (Pi )
is mostly expressed in dB:
Po
10 log (6.61)
Pi
In a medium with absorption this results in
Po
10 log = 10 log e−αz = (10 log e)(−αz). (6.62)
Pi
The attenuation coefficient expressed in dB/m is
The following table presents some important conversions between dB’s and power ratios.
0dB = 1
+1dB ≈ +25% −1dB ≈ −20%
+3dB ≈ +100% of 2× −3dB ≈ −50% or ÷ 2
+6dB ≈ 4× −6dB ≈ ÷4
+10dB ≈ 10× −10dB ≈ ÷10
+20dB ≈ 100× −20dB ≈ ÷100
6–15
In the chapter about lasers we will see that α can be negative, which means that the medium
amplifies the propagating light, instead of absorption!
The parameter β corresponds to the rate by which the phase changes with z and it is called the
propagation constant. The plane wave propagates with phase velocity vp = c/n = ω/(k0 n).
6.7.2 Dispersion
Dispersive media are characterized by a frequency dependent (and wavelength dependent) sus-
ceptibility χ(ν), refractive index n(ν) and speed of light v(ν) = c/n(ν). Optical components such
as prisms and lenses fabricated from dispersive media will refract waves of different wavelengths
into different angles, which leads to chromatic abberation (see section 3.5.1).
Because the speed of light depends on the frequency in a dispersive medium, each frequency
component constituting the wave will experience a different time retardation upon propagation
through the dispersive material. Because of this a short pulse in time will spread out in time. This
effect becomes important upon propagation through kilometers of optical fibers.
The quantity dndλ is called the material dispersion. We noted previously that a monochromatic wave
propagating with propagation constant β has a phase velocity equal to vp = ω/β. However, a
perturbation of the wave, for example by amplitude modulation, travels with another velocity
that is called the group velocity: vg = dωdβ . Correspondingly one defines the group index as N =
dn
c/vg = nef f − λ dλ . For most optical materials the refractive index decreases as the wavelength
increases. Then the group index is larger than the effective index, so the group velocity will be
smaller than the phase velocity. To better understand the concept of group velocity it is instructive
to consider two optical signals with slightly different frequencies, and thus with slightly different
phase velocities (because of material dispersion). The total field shows a beating pattern for the
intensity. This pattern will propagate with a different speed than the two phase velocities.
If a wave is incident on a layered medium - a structure with a number of parallel layers and
interfaces - there are interference effects that are a consequence of the multiple reflections in these
structures. The global reflection and transmission of the structure is dependent on the incidence
angle, the wavelength and the polarization of the incident wave.
The general case of a plane wave incident on a layered medium with N interfaces is treated ele-
gantly by the transfer matrix method. However, this method is beyond the scope of this course.
Here we discuss the simpler case of the three-layer structure (this means one layer in between
two semi-infinite media), as depicted in figure 6.7(a). Such a structure with two parallel semi-
transparent mirrors is a cavity where resonances can exist. It is called a Fabry-Perot etalon. The
discussion is limited to lossless structures, thus with a real refractive index.
As in the case of a single interface, we consider one monochromatic plane wave incident on the
layer structure from a given direction. We assume linear polarization - s-type or p-type - for
6–16
Figure 6.7: (a) Reflection and transmission at a plate. (b) The s-wave and the p-wave.
the E-field, as shown in figure 6.7(b). Any incident field may be considered as a superposition
of such monochromatic linearly polarized plane waves. The following analysis is valid for both
s- and p-polarization. The difference between both situations is contained in the reflection and
transmission coefficients for the interfaces.
One can calculate the global reflection and transmission in two different ways. The first method
closely resembles the physical process, whereas the second method is mathematically more el-
egant. In the first approach the ‘consecutive’ reflections at both interfaces are determined, and
the global reflection and transmission are written as infinite sum series of these contributions. In
the second approach one realizes that every layer contains one forward and one backward plane
wave. By matching the boundary conditions at the interfaces one obtains a linear system that is
easily solvable. Both methods are presented here, and they deliver the same result, of course. For
layered media with more than three layers it is possible to work with both methods, in principle.
However, the first ‘sequential’ method quickly becomes cumbersome, while the second method
remains elegant. For this approach, the system to be solved scales linearly with the number of
layers.
For the first method we consider figure 6.8(a): the plane wave impinges on the first interface, a part
reflects and a part transfers to the second medium. The transmitted part hits the second interface,
with again a partial reflection and transmission. The reflected part goes to the first interface, part
of it goes to medium 1, the other part reflects back etc. etc. All contributions to this sequential
story are indicated as arrows on the figure. However, it has to be clear that each arrow represents
a plane wave that is present in the entire vertical layer. We write the linearly polarized E-field of
the incident field as:
EF,1 (x, z) = Ae−j(kz,1 z+kx,1 x) , (6.64)
6–17
Figure 6.8: Two methods: (a) Sum series of contributions. (b) Global forward and backward plane waves.
with
kz,i = k0 ni cosθi , (6.65)
kx,i = k0 ni sinθi . (6.66)
The index F indicates ‘forward’, thus propagating in the positive z-direction. The total field in
layer 2 is then easily written as the series:
2
−j(kz,2 z+kx,2 x) −j2kz,2 d −j2kz,2 d
EF,2 (x, z) = At12 e 1 + r23 r21 e + r23 r21 e + ... (6.67)
At12
= e−j(kz,2 z+kx,2 x) . (6.68)
1 − r23 r21 e−j2kz,2 d
Here rij (tij ) is the field reflection (transmission) coefficient for incidence from medium i on the
interface with medium j. For the directions of the waves in the three layers one uses Snell’s law:
EB,1 (x, z) = r12 EF,1 (x, 0)e+jkz,1 z + t21 EB,2 (x, 0)e+jkz,1 z (6.73)
t12 t21 r23 e−j2kz,2 d
= A r12 + e−j(−kz,1 z+kx,1 x) . (6.74)
1 − r23 r21 e−j2kz,2 d
6–18
The second method starts from the insight that, upon incidence of a single plane wave, all contri-
butions to the forward field in every layer have the same direction, and thus they form one plane
wave. The same holds for the backward field in each layer. The situation is shown in figure 6.8(b).
In each layer the total forward or backward field is represented by a single plane wave, that we
can write as:
Determining the 4 complex coefficients AB,1 , AF,2 , AB,2 and AF,3 is possible in the following way.
At each interface we write the fields that propagate away from it in function of the fields that
propagate towards it. This amounts to applying the boundary conditions at the interface.
EF,2 (x, 0) = t12 EF,1 (x, 0) + r21 EB,2 (x, 0), (6.80)
EB,1 (x, 0) = r12 EF,1 (x, 0) + t21 EB,2 (x, 0), (6.81)
EB,2 (x, d) = r23 EF,2 (x, d), (6.82)
EF,3 (x, d) = t23 EF,2 (x, d). (6.83)
Solving this system of 4 complex equations with 4 complex unknowns leads to the same result as
with the first method.
For the power reflectance and transmittance of the Fabry-Perot etalon one obtains finally:
The latter expression is only valid if kz,2 is real. It is left to the reader to generalize this expression
for the case where kz,2 is complex. This happens either when the layer is absorptive or when the
field is evanescent in layer 2 due to total internal reflection.
In the next section we examine a symmetrical structure (n1 = n3 ), like the case of a transparent
plate in air. Then, the previous expressions simplify to:
r − r r2 + t t e−j2kz,2 d 2
12 12 12 12 21
R = (6.86)
1 − r122 e−j2kz,2 d
r 1 − e−j2kz,2 d 2
12
= (6.87)
2
1 − r12 e −j2k z,2 d
2
r12 sin2 (kz,2 d) ,
= 4
2 −j2k d
(6.88)
1 − r12 e z,2
2
t12 t21
T =
2 −j2k d
. (6.89)
1−r e 12
z,2
6–19
We can simplify this to:
2
r12 sin2 φ,
R = 4
2 −j2φ
(6.90)
1 − r12 e
2
t12 t21
T = 2 −j2φ
(6.91)
1 − r12 e
with
2π
φ = kz,2 d = n2 dcosθ2 . (6.92)
λ0
For media with a real refractive index we can write the reflectance and transmittance for one
transition, respectively:
R1 = |r12 |2 = r12
2
, (6.93)
T1 = |t12 t21 | = 1 − R1 . (6.94)
T12
T =
1+ R12 − 2R1 cos 2φ
(1 − R1 )2
=
(1 − R1 )2 + 2R1 − 2R1 cos 2φ
1 4R1
= 2 with F = (6.95)
1 + F sin φ (1 − R1 )2
This last equation is also called the Airy equation. Note that we have already calculated this trans-
mission for interference between multiple waves, see section 4.5.2. The maximum transmission is
1, and then we find for perpendicular incidence (cosθ2 = 1) that φ = mπ or d = m 2nλ2 , so the thick-
ness of the layer is an integer times the half wavelength in the material. The minimal transmission
is given by:
1
Tmin = (6.96)
1+F
For sharp maxima we need to ensure that Tmin is as small as possible. Therefore F has to be large,
so R1 needs to be close to 1. In practice, this is difficult because of the available materials.
Figure 6.9 shows the reflectance R for perpendicular incidence on a layer with thickness d and
index n, placed in air. R is presented in function of the wavelength (normalized to nd) for 4
values of n: 1.5, 2, 4 and 8 (this last value is unrealistic for normal materials). One notices that the
reflections drop to 0 if the thickness is an integer times the half wavelength. The reflection dips
become sharper as n increases (and thus R1 increases) and they obtain the character of a resonance.
Such a structure consisting of two semi-transparent mirrors is called a Fabry-Perot resonator.
It is interesting to consider what happens to the reflection or transmission spectrum if the light
incidence is no longer perpendicular but oblique. It is sufficient to realize that the maxima and
minima occur for certain values of φ. If the angle θ increases, then cosθ decreases and thus the
wavelength has to decrease also to keep φ constant. This means that a reflection or transmission spec-
trum always shifts to shorter wavelengths as the light incidence becomes more oblique. This contradicts
the intuition: for oblique angles the light has to travel a longer distance in the layer, and thus one
6–20
Figure 6.9: Reflection of a layer in air with indices n = 1.5 (lower curve), n = 2, n = 4, n = 8 (upper curve).
could expect that the wavelength has to increase to remain in the same maximum or minimum.
We show with figure 6.10 that this reasoning is incorrect: Consider the primary and secondary
contribution to the total transmission. Both contributions are plane waves. To know the phase
difference between them, we have to examine the phase at the same phase front, for example the
plane DD0 . Thus the phase difference is not determined by the path length |BC| + |CD|, but by the
difference in path length between |BC| + |CD| and |BD0 |. This path length difference decreases
as θ increases, while |BC| + |CD| increases! It is an exercise for the reader to show how this path
length difference translates into the phase 2φ.
6.8.2 Reciprocity
Upon careful inspection of equation 6.85 one can see that the power transmission T is invariant
to an exchange of layer 1 by layer 3 and vice versa. More precisely: the power transmission
is identical for transmission from left to right at an angle of incidence θ1 and from right to left
at an angle of incidence θ3 (connected to θ1 by Snell’s law). This remarkable property is called
6–21
Figure 6.11: Reciprocity in an N -layer slab
6.8.3 Coatings
Layer structures can be employed to increase or decrease the reflection of a surface. This is useful
for the design of anti-reflection coatings (AR-coatings) for lenses, and for the design of dielectric
mirrors. Most of the time one uses perpendicular incidence.
In designing an AR-coating we ensure that the reflection at the front of the film interferes destruc-
tively with the reflection at the back of the film. If n1 < n2 < n3 then one needs: (note: extra phase
shift π for reflection at interface 1-2 and interface 2-3!)
1 λ0 λ2
d= = , (6.97)
4 n2 4
hence the name quarter-wave layer. This is illustrated in figure 6.12 for the first two contributions
of the reflected field. In practice these are the most important contributions (note that our analysis
does take all reflections into account).
6–22
Combining the previous equation with the Fresnel coefficients for perpendicular incidence:
ni − nj
rij = (6.98)
ni + nj
2ni
tij = (6.99)
ni + nj
Figure 6.12: AR-coating consisting of one layer. (a) Principle, (b) reflection
√ spectrum for an AR-coating
designed for the telecom wavelength of 1550nm. n1 = 1, n3 = 3.2, n2 = 3.2 = 1.79, d = 217nm.
A HR-coating could consist of a quarter-wave layer made from a higher index than both consid-
ered media. In practice this is often not realizable, therefore one employs a periodic structure
of quarter-wave layers alternating between high and low index, see figure 6.13. Together they
behave as a Bragg reflector.
If the thickness of the consecutive layer can be controlled so that:
λ0
nH dH = nL dL = (6.101)
4
6–23
then the reflected beams from the different interfaces will all interfere constructively, leading to a
large reflection coefficient. Using the matrix method one can obtain for R:
2N 2
nH
1 − nL
R= 2N . (6.102)
nH
1 + nL
nH
R converges to 1 as N increases. The convergence improves as the ratio nL becomes larger.
A HR-coating consisting of silver sulfide (nH = 2.32) and magnesium fluoride (nL = 1.38) has
a reflection of 98.9% already after 13 layers, at λ = 633nm. Such a highly-reflective mirror is
used for fabrication of a helium-neon laser cavity.
Figure 6.13: HR coating. (a) Principle. (b) Reflection of a coating for a He-Ne laser at wavelength λ =633nm.
nH =2.32 (ZnS), nL =1.38 (MgF2 )
Exercise: Explain why in the AR-coating a quarter wavelength thick layer leads to destructive
interference for the reflected light, whereas in the HR-coating quarter wavelength thick layers
lead to constructive interference for the reflected light.
For more complicated applications (broadband and narrowband filters, power and polarization
splitters. . . ) one uses specialized CAD-software.
Figure 6.14 shows an example of a design for sunglasses. The demands were the following:
The designed coating has 29 layers of SiO2 and TiO2 with thicknesses between 20nm and
200nm on a glass substrate.
6–24
Figure 6.14: Transmission of sunglasses (
c
1995-98 Software Spectra, Inc., http://www.sspectra.com/).
6.9 Scattering
The scattering of light can be seen as the deviation from a straight trajectory when the electro-
magnetic (EM) wave (light) encounters obstacles or non-uniformities in the medium in which it
travels. The scattering mechanisms that we will discuss here involve scattering particles which
can be assumed spherical. When the EM wave encounters a particle it will cause a periodic per-
turbation in the electron orbits within the molecules of the particle. This perturbation has the
same frequency as the incoming EM-wave. The separation of the charges in the molecule due to
the perturbation is called the induced dipole moment. This oscillating dipole moment is now a
new EM source, resulting in scattered light.
When the wavelength of the scattered light is the same as the wavelength of the incoming wave,
we say that the scattering is elastic. This means that no energy is lost in the scattering process.
When energy is partly converted (e.g. to heat or vibrational energy) and the resulting wavelength
is larger than the original wavelength, the scattering process is said to be inelastic. Examples of
such inelastic scattering are Brillouin and Raman scattering. We will now discuss two elastic light
scattering mechanisms: Rayleigh scattering and Mie scattering.
Rayleigh scattering (named after Lord Rayleigh) is caused by particles smaller than the wave-
length of the incident light. It can occur in solids or liquids but it is mostly seen in gasses. The
criterion for Rayleigh scattering is: α<<1, with
2πr
α= . (6.103)
λ
r is the radius of the particle and λ is the wavelength of the incident light. It can be shown that
in the Rayleigh regime, shorter wavelengths are scattered more efficiently (scaling1/λ4 ). This
explaines why the daytime sky looks blue. The (shorter) blue wavelengths are redirected more
efficiently towards earth than the (longer) red ones.
Mie scattering (named after Gustav Mie) is the general scattering theory without limitations on
the particle size. For large particles, this theory converges to geometric optics. It can also be
used for very small particles but in that case the Rayleigh theory is preferred due to the simplicity
compared to the Mie theory. E.g. Mie scattering explains why clouds are white as it involves
scattering of sunlight from particles (in this case water droplets) which are small but larger than
the wavelength of the light. Other examples are scattering from dust, smoke, pollen,...
6–25
Bibliography
[ST91] B.E.A. Saleh and M.V. Teich. Fundamentals of Photonics. John Wiley and Sons, ISBN 0-471-
83965-5, New York, 1991.
6–26