Waves in Media: Ashcroft and Mermin, Solid State Physics (Saunders College, 1976, Page 553)
Waves in Media: Ashcroft and Mermin, Solid State Physics (Saunders College, 1976, Page 553)
Waves in Media
2.1 Introduction The propagation of waves in a medium depends on the magnetic permeability and electric
permittivity functions µ (ω) and ε (ω) for the medium. We will generally deal with systems in which µ (ω) can be
approximated as having the value 1. A general form often used to approximate the electric permittivity is the Sellmeier
equation
X Am
ε (ω) = 1 + . (2.122)
m
ω 2m − iγ m ω − ω 2
This equation is based on the assumption that the system behaves as a set of resonances. This will be observed in the two
systems we will examine in this chapter. First we will consider electromagnetic waves in ionic crystals. The discussion will
be restricted to frequencies below that of the electronic interband transitions. In this frequency range the dielectric function
is determined by the ionic displacements and the polarizabilities of the ions. The second system will be electrical conductors,
either metals or plasmas. In these systems the frequency dependent conductivity determines the dielectric function.
The analysis of the models will be followed by a discussion of wave packets in dispersive materials. This discussion
will use the permittivities obtained for the two models. The propagation of electromagnetic waves in a medium involves the
index of refraction of the medium. From the deduced analytic properties of the complex index of refraction we will obtain
the KramerKronig relationship between the real and imaginary parts of the index of refraction.
We will close the chapter with a discussion of reectance and transmittance of electromagnetic waves at the interface
between media.
2.2 Dielectric constant for ionic crystals Ionic crystals such as the alkalihalides are cubic crystals. Note
that electromagnetic fields only interact with the ‘optic modes’ for the crystal. These are the modes in which the positive and
negative ions move in opposite directions. The dielectric function of cubic crystals is independent of the direction of the
electric field and can be characterized by three parameters, the static dielectric constant εs , the optical dielectric constant
εopt , and a characteristic resonant frequency ω T . For reference typical resonant frequencies are in the range 0.02eV 0.04eV,
optical frequencies are in the range 2eV 3eV, and the electronic interband transitions begin around 4 or 5eV. An empirical
dielectric function for these crystals is
εopt − εs
ε (ω) = εopt + ω 2T (2.123)
ω2 + iγω − ω 2T
We will find that ω T can be identified as the transverse optical phonon frequency for the crystal. A damping factor has been
included to indicate what a real dielectric function might look like. This is an approximation to the Sellmeier equation, given
in Eq.2.122, in a frequency range near one resonant term and far from any other resonance. Figure 2. shows a typical
dielectric function (KBr1 : εs = 4.90, εopt = 2.34, and ω T = 2.26 × 1013 rad/s, 14.3 meV ) with no damping.
1 Ashcroft and Mermin, Solid State Physics (Saunders College, 1976, page 553)
32
Real ε(ω) for KBr
ε(ω)
20.0
10.0
ωL/ω
T
0.0
-10.0
-20.0
0 1 2 3 4 5
ω / ωT
.Figure 2.
2.2.1 Longitudinal optical crystal vibrations From the permittivity we can obtain the frequency for a
longitudinal optical mode of crystal vibration. These modes are characterized by a longitudinal charge density wave (that is,
a charge density disturbance which propagates along the direction of the wave). This wave would generate a longitudinal
electric field wave when it generates a current density. The divergence of the current density gives the time rate of change of
the charge density. If this is selfsustaining in the absence of externally applied fields it is a normal mode of oscillation of the
system. The electric field set up by the microscopic charge density will satisfy ∇ · E (r, ω) = 4πρm (r, ω) . In the absence
of an ‘external charge density’ we must have that ∇ · D (r, ω) = 0. Since D (r, ω) = ε (ω) E (r, ω) and ∇ · E (r, ω) 6= 0 we
must find that ε (ω L ) = 0 where ω L is the frequency of the longitudinal dynamic mode. The solution is given by
r
εs
ωL = ωT (2.124)
εopt
(For reference2 : KBr ω L = 3.14 × 1013 rad/s and (ω L /ω T )2 = 1. 9) This is known as the LyddaneSachsTeller relation.
In the first approximation the longitudinal optical modes of vibration of a crystal have a constant frequency independent of
the wavelength of the charge density wave. The location of this frequency is noted in Fig. 2. .
2.2.2 Transverse electromagnetic waves in the crystal The transverse optical modes do not generate a
charge density. They do involve current densities but the electromagnetic fields generated by these currents do not contribute
an appreciable amount to the forces involved in the crystal vibrations. However an electromagnetic wave propagating
through the material can couple to the transverse optical mode of vibration.
2 Woods, Brockhouse, Cowley, and Cochran, Phys. Rev. 131 1027 (1963)
33
.Figure 3.
2π
These modes have approximately constant frequency and wave vectors, k = , ranging from zero to of order
λT O
1 rad/nm. That is, λT O ≥ 2π nm. Examples of the dispersion curves for longitudinal and transverse optical and
acoustic phonon modes are shown in Fig.3 . Compare the range of phonon wave vectors to a typical wave vector for an
ω 1013 rad/s
ω = 1013 rad/s electromagnetic wave: k = = ≈ 3 × 10−5 rad/nm. In the ω versus k curves shown
c 3x108 · 109 nm/s
in Fig. 3 the ω versus k of an electromagnetic wave would lie on the right hand vertical axis. To search for transverse
modes we consider the propagation of a transverse electromagnetic wave in the crystal. In the following we assume for
convenience that µ (ω) = 1. Since we seek transverse modes we only consider electric fields satisfying ∇ · E (r, ω) = 0.
[−iω]2
−∇2 E (r, ω) = −ε (ω) E (r, ω) (2.125)
c2
ω2
= ε (ω) E (r, ω)
c2
or, in wave vector space,
ω2
k2 E (k, ω) = ε (ω) E (k, ω) (2.126)
c2
34
so that
ω2
(k2 − ε (ω)
)E (k, ω) = 0
c2
Thus, for a transverse electromagnetic wave to propagate in the medium the wave vector and angular frequency must satisfy
the dispersion relation
ω2
k2 = ε (ω) (2.127)
c2
or, using the dielectric function without damping,
εopt − εs εopt ω 2 − ω 2T εs
ε (ω) = εopt + ω 2T 2 =
2
ω − ωT ω 2 − ω 2T
kc ω2 T T
= ε (ω) 2 = ω2
ωT ωT ω 2 − 1
T
ω2
or, letting x = ω 2T
,
µ ¶2 µ ¶2
2 kc kc
εopt x − (εs + )x + =0
ωT ωT
³ ´2 ³ ´2 ³ ´2 1/2
2 (εs + kc
) (εs + ωkc )2 − 4εopt ωkc
ω ωT T T
= ± (2.128)
ω 2T 2εopt
4ε2opt
³ ´2 ³ ´2 1/2
(εs + kc
)
4εopt ωkcT
ω2 ωT
1 ± 1 −
= ³ ´
ω 2T 2εopt
kc
2
2
(εs + ωT )
³ ´2 ³ ´2
(εs + ωkcT ) 4ε
1 opt ωT
kc
kc
≈ 1 ± 1 ∓ 2 when −→ 0
2εopt 2 εs ωT
³ ´2 ³ ´2
kc kc
(εs + ωT ) 1 4ε opt ωT kc
≈ 1 ± 1 ∓ ³ ´2 when −→ ∞
2εopt 2 ω T
(εs + ωkcT )2
As kc
ωT −→ 0
³ ´2
kc
ω2 (εs + ωT ) εs
≈ 2 −→
ω 2T upper branch (+) 2εopt εopt
³ ´2 ³ ´2
ω2 (εs + ωkcT ) 1 4εopt kc
ωT
≈ −→ 0
ω 2T lower branch () 2εopt 2 ε2s
As kc
ωT −→ ∞
35
³ ´2 ³ ´2
kc kc
ω 2 (εs + ωT )
εs ωT
≈ 2 −→ +
ω 2T upper branch (+) 2εopt εopt εopt
³ ´2
ω2 2εopt ωkcT
≈ ³ ´2 −→ 1
ω 2T lower branch ()
2εopt (εs + ωkcT )
These two dispersion relations for the electromagnetic waves are shown in Fig. 4 with the minus sign giving the lower
branch and the
q plus sign the upper. We note that an electromagnetic wave qin the frequency range ω T < ω < ω L ( or
ω
1< < εεopt s
= 1.39) can not propagate in the crystal, where ω L = εεopt
s
ω T . The transverse optical modes of the
ωT
crystal will lie in this range of frequencies.
3.00
2.00
ω = kc/ εs1/2
1.00
0.00
0 1 2 3 4 5
kc / ωΤ
.Figure 4.
The lower branch does not appear to be photonlike (maybe phononlike?). In contrast the dispersion relation for the
√
upper branch approaches ω = kc/ εopt for large k, exhibiting the standard photonlike behavior in the range of optical
frequencies. This behavior of the dispersion relation for the electromagnetic wave implies that ω T is a frequency for a
transverse vibrational mode for the crystal. A wave entering the crystal with a frequency near the resonant frequency of the
transverse optical mode of the crystal would be strongly coupled to the crystal mode. The result would be a propagating
wave with a mixture of the characteristics of the mechanical and electromagnetic waves. The gap in the range of frequencies
of propagating waves is due to this coupling. The transverse optical vibration at ω T depresses the frequency of the
electromagnetic wave. The long wavelength electromagnetic wave oscillating at the frequency of the longitudinal optical
36
vibration would provide the electric field that is generated in the longitudinal mode. The ω L therefore will provide the other
end of the frequency band for the ‘stongly coupled’ mechanicalelectromagnetic mode.
Finally, we note that the frequency difference between the longitudinal and transverse optical modes, ω L − ω T , can be
attributed to the electric field which is generated by the longitudinal mode. Even in the long wavelength region this provides
an extra restoring force for the displaced ions thereby making the longitudinal mode frequency greater than that of the
transverse mode.
2.2.3 Index of refraction for an ionic crystal Generally, when dealing with the propagation of an
electromagnetic wave in a medium, the frequency of the wave is determined by the boundary conditions and sources.
Index of Refraction
12
0
0 0.5 1 1.5 2
Normalized frequency
real index
imaginary index
n=1
.Figure 5.
The relation between the wave vector and the angular frequency is given by k = kr + iki with
ω
kr (ω) = n (ω) = Re k (2.129)
c
ω
ki (ω) = κ (ω) = Im k (2.130)
c
From Eq. 2.127 we find
ω2
k2 = [Re ε (ω) + i Im ε (ω)] (2.131)
c2
ω2
= [n (ω) + iκ (ω)]2
c2
ω2
= [n2 (ω) − κ2 (ω) + 2in (ω) κ (ω)]
c2
1
n2 (ω) = [Re (ε (ω)) + |ε (ω)|] , n (ω) > 0 (2.133)
2
1
κ2 (ω) = [− Re (ε (ω)) + |ε (ω)|] , sgn [κ (ω)] = sgn [Im ε (ω)] (2.134)
2
For an ionic crystal
εopt − εs ω 2 − iγω − ω 2T
ε (ω) = εopt + ω 2T ·
ω 2 + iγω − ω 2T ω 2 − iγω − ω 2T
εopt − εs
= εopt + ω 2T 2 [ω 2 − ω 2T − iγω]
[ω − ω 2T ]2 + γ 2 ω 2
εopt − εs
Re ε (ω) = εopt + ω 2T [ω 2 − ω 2T ] (2.135)
[ω 2 − ω 2T ]2 + γ 2 ω 2
ω2
εopt − εopt 2L
ωT
= εopt + ω 2T 2 2 [ω 2 − ω 2T ]
[ω − ω T ] + γ 2 ω 2
2
[ω 2 − ω 2T ]2 + γ 2 ω 2 + (ω 2T − ω 2L )[ω 2 − ω 2T ]
= εopt
[ω 2 − ω 2T ]2 + γ 2 ω 2
[ω 2 − ω 2T ][ω 2 − ω 2T + ω 2T − ω 2L ] + γ 2 ω 2
= εopt
[ω 2 − ω 2T ]2 + γ 2 ω 2
[ω 2 − ω 2T ][ω 2 − ω 2L ] + γ 2 ω 2
= εopt
[ω 2 − ω 2T ]2 + γ 2 ω 2
¡ 2 2
¢
(εopt − εs ) γω 2 εopt ω T − ω L γω
Im ε (ω) = −ω T 4
2 = −ω T 2 (2.136)
(ω 2 − ω 2T ) + γ 2 ω 2 (ω 2 − ω 2T ) + γ 2 ω 2
¡ 2 ¢ 2
ω − ω 2L + γ 2 ω 2
|ε (ω)| = εopt 2 (2.137)
(ω 2 − ω 2T ) + γ 2 ω 2
The real and imaginary parts of the index of refraction for KBr are shown in Fig. 5 .
The electric field for a plane electromagnetic wave travelling in the +z direction is given by
Z ∞
E (z, t) = A (ω) exp [ikz − ωt] dω
−∞
Z ∞ h ω i
= A (ω) exp i (n (ω) + iκ (ω))z − ωt dω
−∞ c
Z ∞ h ω i h ω i
E (z, t) = A (ω) exp i (n (ω) z − ct) exp − κ (ω) z dω, A3 (ω) = 0 (2.138)
−∞ c c
In general εs > εopt and, therefore, ωκ (ω) ≥ 0 for all ω. It follows that Eq. 2.138 describes a damped plane wave. The
properties of this wave will be analyzed in a later section.
.................................................................................................................
.................................................................................................................
38
Section 2.3 Electronic plasmas
2
(a) Calculate the wave number spectrum |A (k)| for each of the above forms for f (x).
(b) Explicitly evaluate the rms deviations from the mean, ∆x and ∆k.
(c) Using graphs compare the functions |f (x)|2 and the functions |A (k)|
2
. Include two graphs one comparing the functions f (x) and the other comparing the functions |A (k)| .
Assignment 6b: Jackson Problem 7.20 A homogeneous, isotropic, nonpermeable dielectric is characterized by a
complex index of refraction n (ω) . (a) Show that the general solution for plane waves in one dimension can be written
Z ∞ h i
1
u (x, t) = √ e−iωt A (ω) eiω n(ω)x/c + B (ω) e−iω n(ω)x/c dω
2π −∞
where u (x, t) is a component of E or B. (b) If u (x, t) is real show that n (−ω) = n (ω)∗ . (c) Show that, if u (0, t) and
[∂u (x, t) ∂x]x=0 are tha boundary values of u (x, t) at x = 0, the coefficients A (ω) and B (ω) are
µ ¶ Z ∞ · µ ¶ ¸
A (ω) 1 1 iωt ic ∂u (x, t)
= √ e u (0, t) ∓ dt
B (ω) 2 2π −∞ ωn (ω) ∂x x=0
.................................................................................................................
.................................................................................................................
2.3 Electronic plasmas A plasma is an electrically charged gas consisting of ions and electrons. It is similar to a
liquid conductor, but is generally distinguished from a liquid conductor by density. A liquid conductor has a number density
of the order of 1022 cm−3 whereas a plasma has a number density of the order of 1018 cm−3 . A plasma can be thought of as
a gaseous conductor. [ See S. Gartenhaus, Elements of Plasma Physics, Holt, Rinehart and Winston, New York 1964 for an
introduction to plasmas.]
We consider here a system composed of ‘free electrons’ and a uniform background of positive charge to provide charge
neutrality for the system. In most cases the radiation emitted (lost) by the accelerated ions in the plasma is negligible, as
are quantum effects (which are relevant only at high densities and low temperatures.) In a gas of density 1018 cm−3 the
average separation distance is ≈ [10−18 ]1/3 =10−6 cm = 100 Å. The ”macroscopic” treatment of electric and magnetic
fields in a plasma assumes that properties are averaged over distances ≥ 100 Å. The electrons in the gas can also undergo
close collisions with the ions and with other electrons. The collisions between electrons thermalize the electron ‘gas’ while
the collisions with ions will reduce the momentum of the electron gas.
Our analysis will use the linearized Boltzmann equation for an electron density in phase space, ρ (r, v, t) (with units
cm−3R(cm/ sec)−3 ). The equilibrium
R electron density will be velocity dependent but not spatially dependent, i.e., ρ0 (v)
with ρ0 (v) d3 v = ρ0 and vρ0 (v) d3 v = 0. (The ρ0 has units cm−3 .) The linearized Boltzmann equation assumes
that the total time derivative of ρ (r, v,t) taken along a particle trajectory is proportional to ρ (r, v,t) − ρ0 (v), the deviation
of the electron density from equilibrium. The proportionality constant is −1/τ and is due to ‘collisions’ not included in the
dynamical equations for the system. In the model considered here the excluded interactions are close collisions between
electrons and collisions with the ions.
39
Section 2.3 Electronic plasmas
As an example of a collision term we consider the electronelectron collisions3 . The collision crosssection is a function
of the relative speed of the particles and the scattering angle. We note that the center of mass motion of the particles
(conservation of momentum) and their relative speed (conservation of energy) are not changed by the collision. The collision
process transfers particles between the velocity pairs {v, v0 } and {u, u0 } and can be classified by v0 and, taking v − v0 as
the z axis, the angle (θ, φ) for u − u0 .
· ¸ ZZ
dρ (r, v,t)
= σ (|v − v0 | ; θ, φ) |v − v0 | ρ (r, u,t) ρ (r, u0 , t) d3 v 0 dΩ
dt collision
ZZ
− σ (|v − v0 | ; θ, φ) |v − v0 | ρ (r, v,t) ρ (r, v0 , t) d3 v 0 dΩ (2.139)
If ρ(r, v,t) is the Boltzmann distribution ρ0 (r, v,t) , the collision term is zero. In the case that
is ‘small’ then the collision term can be approximated by a linear function of δρ (r, v,t) (essentially using the principle of
detailed balance) to obtain
· ¸
dρ (r, v,t)
dt collision
ZZ
= σ (|v − v0 | ; θ, φ) |v − v0 | [δρ (r, u,t) ρ0 (r, u0 , t) + ρ0 (r, u,t) δρ (r, u0 , t)] d3 v 0 dΩ
ZZ
− σ (|v − v0 | ; θ, φ) |v − v0 | [δρ (r, v,t) ρ0 (r, v0 , t) + ρ0 (r, v,t) δρ (r, v0 , t)] d3 v 0 dΩ (2.140)
δρ (r, v,t)
∼ − (2.141)
τ
This will provide a reasonable, semiempirical expression for the relaxation of the distribution to equilibrium. The Boltzmann
equation itself is a hybrid combination of dynamics and probability.
The total time derivative along a trajectory will involve the time dependence of r and v on the trajectory.
d ∂ ρ (r, v,t) − ρ0 (v)
ρ (r, v,t) = ρ (r, v,t) + v · ∇r ρ (r, v,t) + a · ∇v ρ (r, v,t) = − . (2.142)
dt ∂t τ
The acceleration of the electrons is due to the electric field generated by the spatial variation in the electron density and by
any externally generated electric and magnetic fields. In particular the latter will be the transverse electric and magnetic
fields of an electromagnetic wave. The acceleration of an electron located at r and moving with velocity v at time t is
e h v i
a=− E (r, t) + × B (r, t) (2.143)
m c
The fields satisfy Maxwell’s equations:
Z
∇ · E (r, t) = ε−1
i (−4πe) [ρ (r, v,t) − ρ0 (v)] d3 v, (2.144)
1 ∂
∇ × E (r, t) = − B (r, t) , (2.145)
c ∂t Z
εi ∂ 4πe
∇ × B (r, t) = E (r, t) − vρ (r, v,t) d3 v, (2.146)
c ∂t c
∇ · B (r, t) = 0. (2.147)
3 Kerson Huang, Statistic Mechanics (Problem of Kinetic Theory, Chapt. 3, Wiley, 1963) See also Section 5.4
The quantum mechanical system is covered by Pines, The Many Body Problem (Benjamin, 1961) An interesting approach is provided by an article by
Ehrenreich and Cohen (Phys, Rev., 115, 786790 (1959)) which is reprinted on pp 255259.
40
Section 2.3 Electronic plasmas
Neglecting the magnetic force in our transport equation we find the δρ must satisfy
∂ e δρ (r, v,t)
δρ (r, v,t) + v · ∇r δρ (r, v,t) − E0 exp [i (k · r − ωt)] ·∇v ρ0 (v) = − (2.153)
∂t m τ
We look for a electron density variation of the form
The last term on the right is zero since both integrations over φ vanish. We are then left with
ZZZ
E0 · k eτ k·v dρ0 (v) 3
n (r, t) = i 2 exp (ik · r − iωt) d v (2.159)
k mv ωτ − v · kτ + i dv
¡ £ ¤¢
It suffices for our analysis to carry out an expansion of the integrand in powers of v · k / ω + iτ −1
ZZZ · ¸
E0 · k i(k·r−ωt) e k·v k·v dρ0 (v) 3
n (r, t) = i 2 e −1
1+ −1
+ d v (2.160)
k mv ω + iτ ω + iτ dv
ZZZ · ¸n
E0 · k e k·v dρ0 (v) 3
= i 2 ei(k·r−ωt) Σn=1,∞ −1
d v
k mv ω + iτ dv
Z ∞ · ¸n Z 1
E0 · k 2πe kv dρ0 (v)
= i 2 ei(k·r−ωt) Σn=1,∞ −1
vdv cosn θd(cos θ)
k 0 m ω + iτ dv −1
Z ∞ µ ¶j
E0 · k i(k·r−ωt) 2πe kv 2 dρ0 (v)
= i 2 e Σj=2,4,6,... vdv
k 0 m ω + iτ −1 j + 1 dv
Z µ ¶2 " µ ¶2 #
E0 · k i(k·r−ωt) ∞ 4πe kv 3 kv dρ0 (v)
= i 2 e −1
1+ −1
+ ... vdv
k 0 3m ω + iτ 5 ω + iτ dv
Z ∞
4πe dρ0 (v) 3
iE0 · kei(k·r−ωt) −1 2
v dv (2.161)
3m(ω + iτ ) 0 dv
· Z ∞ ¸
4πe
= iE0 · kei(k·r−ωt) v 3
ρ0 (v) |∞
0 − 3 ρ0 (v) v 2
dv
3m(ω + iτ −1 )2 0
ZZZ
e
= −iE0 · kei(k·r−ωt) ρ0 (v) d3 v
m(ω + iτ −1 )2
Using this first term as an approximation for n (r, t), one obtains
eρ0 E0 · k
n (r, t) = −i ei(k·r−ωt) (2.162)
m (ω + iτ −1 )2
independent of the form taken for the equilibrium electron density. We should note that n (r, t) = 0 if E0 · k =0.
42
Section 2.3 Electronic plasmas
Z 2π Z π Z 2π Z π
0
cosn θvx i vyj sin θdθdφ = v 2 cosn θ sini+j θ cosi φ sinj φ sin θdθdφ (2.165)
0 0 0 0
Z 1 Z 2π
0
= v2 cosn θ sini+j θd(cos θ) cosi φ sinj φdφ; u = cos θ
−1 0
Z 1
i+j Z 2π
n0
= v2 2
u [1 − u ] 2 du cosi φ sinj φdφ; (2.166)
−1 0
= 0 if n is odd (This applies to all terms along x̂ and ŷ)
0
Z cos π Z 2π
= v2 cos2 θ sini+j θd(− cos θ) cosi φ sinj φdφ if n=2 ( ẑ terms)
(2.167)
cos 0 0
So all the terms with along x̂ and ŷ vanish. The terms along ẑ vanish if i = j = 1 because
Z 2π
cos φ sin φdφ = 0 (2.168)
0
Some of the terms for i = j = 2 do not vanish. A non zero term comes from the third term in the expansion.
43
Section 2.3 Electronic plasmas
If τ is very large (the relaxation to equilibrium takes a long time compared to the period of the field) then the electron
current density will lead the electric field in time by 90◦ . That is, using −i = e−iπ/2
e ρ0
je (r, t) ≈ e−iπ/2 E0 exp (ik · r − iωt) (2.171)
m ω + iτ −1
e ρ0
= E0 e−iπ/2+ik·r−iωt
m ω + iτ −1
e ρ i τ f ield
= E0 0 [1 − + ...]e−iπ/2+ik·r−iωt
m ω 2π τ
e ρ e ρ
≈ = E0 0 ei[k·r−ωt−π/2] = E 0 e−iπ/2
m ω m ω
π
This means that the electric field has the same phase as the current density at a later time t0 = t + . Physically the current
2ω
density lags the force by 90◦ as expected.
e2 ρ0 τ
jc = −eje = E0 (2.173)
m
e2 ρ0 τ
σ0 = . (2.174)
m
. ................................................................................................................
.................................................................................................................
Assignment 7: For a metal the electron density would be approximated by the ‘fermi sea’,
µ ¶3
m∗ ¡ ¢1/3 }
ρ0 (v) = 2 Θ (vf − v) , vf = 3π2 ρ0 (2.175)
h m∗
with m∗ the effective mass of the electron. The equilibrium electron density for a charged plasma at temperature T is given
by the Boltzmann distribution
µ ¶3/2 µ ¶
m mv 2
ρ0 (v) = ρ0 exp − (2.176)
2πkB T 2kB T
Using these equilibrium densities evaluate the correction terms for the electron density, n (r, t) , and electron current
density, j (r, t) .
.................................................................................................................
.................................................................................................................
45
Section 2.3 Electronic plasmas
2.3.1 Longitudinal plasma oscillation The electron density given by Eq. 2.162 will create a charge density
which generates an electric field. We seek the condition for which the electric field generated by this charge density equals
the electric field which creates the charge density variation. In this case the field and charge density are related by Gauss’s
law (εi is the permittivity of the background ions)
e2 ρ0
4π =1 (2.178)
mεi (ω + iτ −1 )2
−∂
∇ × (∇ × E) = ∇ × [ B] (2.180)
c∂t
−∂
−∇2 E + ∇(∇ · E) = (∇ × B) (2.181)
c∂t
−∂ ∂D 4π
= ( − Jo ) (2.182)
c∂t c∂t c
Note that for a transverse wave ∇ · E = ik · E = 0 and (using the first order term for Jo = eje )
−1 ∂ 2 4π ∂ e2 ρ0
−∇2 E = D+ [−i E0 exp (ik · r − iωt) ] (2.183)
c2 ∂t2 c c∂t m ω + iτ −1
ω2 4πω e2 ρ0
k 2 E0 = εi 2 E0 − 2 E0 (2.184)
c c m ω + iτ −1
Hence, the dispersion relation (see Eq.127) emerges from:
(ck)2 4π e2 ρ0
E0 = [εi − ]E0 (2.185)
ω2 ω m ω + iτ −1
= ε(ω)E0 (2.186)
46
Section 2.3 Electronic plasmas
4πe2 ρ0
ε (ω) = εi − . (2.187)
m ω (ω + iτ −1 )
4πe2 ρ0
Using ω 2p = mεi from Eq.2.179
" # " #
ω 2p ω 2p τ 2 (ωτ − i)
ε (ω) = εi 1− = εi 1 − (2.188)
ω (ω + iτ −1 ) ωτ (ω 2 τ 2 + 1)
4πσ 0 4πσ 0 (ωτ − i)
= εi − = εi − (2.189)
ω (ωτ + i) ω (ω 2 τ 2 + 1)
4πσ 0 τ 4πσ 0
= εi − 2 2 +i (2.190)
(ω τ + 1) ω (ω 2 τ 2 + 1)
" #
ω 2p τ 2 εi ω 2p τ 2
= εi 1 − 2 2 +i (2.191)
(ω τ + 1) ωτ (ω 2 τ 2 + 1)
showing the explicit relationships between the permittivity, the plasma frequency, and the conductivity. Figure 6 shows
the real and imaginary parts of the permittivity for a metal with ω p τ = 10. We note that the real part of the permittivity is
negative below the plasma frequency and monotonically approaches permittivity of the background ions above the plasma
frequency.
0.00
-1.00
ωτ = 10
real ε
-2.00
-3.00
5.0 10.0 15.0 20.0 25.0 30.0
ωτ
.Figure 6. Real and imaginary parts of the electrical permittivity for a metal with ω p τ = 10.
Equations 2.133 and 2.134 can be used to evaluate the real and imaginary parts of the index of refraction. The real and
imaginary parts of the index of refraction are shown plotted versus ωτ in Fig. 7 . In this plot the product, ω p τ , of the plasma
frequency and the relaxation time, τ , has been taken to be 10 and the permittivity of the background ions has been taken to
be 1
47
Section 2.4 Propagation of plane waves in a medium
0
0 5 10 15 20
real index
imaginary index
n=1
.Figure 7
We note that in the region near zero frequency the imaginary part of the permittivity is very large (approaching infinity
at zero frequency). In this region the real and imaginary parts of the index of refraction are nearly equal. Indeed the real
part of the index of refraction lies below the imaginary part for all frequencies below the plasma frequency. Note that the
energy dissipated is proportional to the imaginary part of the permittivity (see Eq. 111 in Chapter 1). It will therefore not
be meaningful to speak of a wave propagating in the medium at frequencies below the plasma frequency.
2.4 Propagation of plane waves in a medium We now have two models for the index of refraction. Our
next task is to investigate the propagation of a plane electromagnetic wave in an ionic crystal and in a metal. Reviewing the
permittivities obtained for the models we note that the real part of the permittivity is an even function of ω and the imaginary
part is an odd function of ω. The properties of the real and imaginary parts of the indices of refraction are obtained from
those of the permittivities and Eqs. 2.133 and 2.134. Like the permittivity we find that the real part of the index of refraction
is an even function of ω and the imaginary part is an odd function of ω. Using these properties a plane polarized wave
travelling in the +z direction can be described by the expression
Z ∞
E (z, t) = E0 (ω) exp(i[Re k(ω) + i Im k(ω)]z − ωt) dω + c.c.
0
Z ∞
ω ω
= E0 (ω) exp(i (n (ω) z − ct) exp(− κ (ω) z) dω + c.c. (2.192)
0 c c
We will restrict our analysis to the case in which |E0 (ω)| has a single, ‘narrow’ peak of width ∆ω at ω 0 . The peak will be
considered narrow if
∆ω dn (ω 0 )
¿ 1.
n (ω 0 ) dω 0
2.4.1 Group velocity For a wave composed of a narrow range of frequencies the wave vector k (ω) can be
expanded about the central frequency to obtain an approximate expression for the propagation of the wave. In our example
48
Section 2.4 Propagation of plane waves in a medium
· ¸ · ¸ 2
ω ω0 dn (ω 0 ) ω − ω 0 1 dn (ω 0 ) d2 n (ω 0 ) (ω − ω 0 )
n (ω) = n (ω 0 ) + n (ω 0 ) + ω 0 + 2 + ω0 + ... (2.193)
c c dω 0 c 2 dω 0 dω 20 c
A similar expression holds for ωκ (ω) /c. Working to first order in ω − ω 0 we find
ω0 ω0
E (z, t) = exp(i (n (ω 0 ) z − ct) exp(− κ (ω 0 ) z)
Z ∞ c µ ½·c ¸ ¾¶
ω − ω0 dn (ω 0 )
× E0 (ω) exp i n (ω 0 ) + ω 0 z − ct dω + c.c. (2.194)
0 c dω 0
where we have neglected the variation in the imaginary part of the index of refraction. In this approximation the wave travels
with the group velocity
· ¸−1
dn (ω 0 )
vg (ω 0 ) = c n (ω 0 ) + ω 0 (2.195)
dω 0
i.e.
Z ∞ µ ¶
ω − ω0
E (z, t) = E0 (ω) exp i {z − vg (ω 0 ) t} dω
0 vg (ω 0 )
ω0 ω0
· exp(i (n (ω 0 ) z − ct) exp(− κ (ω 0 ) z) + c.c.
c c
ω0 ω0
= F(z − vg (ω 0 ) t) exp(i (n (ω 0 ) z − ct) exp(− κ (ω 0 ) z) + c.c. (2.196)
c c
This should describe a ‘broad’ wave packet with the envelope function travelling with speed vg (ω 0 ) and the underlying
wave travelling with speed c/n (ω 0 ) . Figure 8 shows the group velocity for the index of refraction of an ionic crystal shown
in Fig. 5 . The normalized frequency is ω/ω T with the approximation (for the peaked E0 (ω)) that ω ≈ ω 0 . The group
velocity approaches 0 at ω/ω T = 1 where n (ω T ) diverges.
0.25
0
0 0.25 0.5 0.75 1
normalized frequency
49
Section 2.4 Propagation of plane waves in a medium
Fig. 9 shows the group velocity, above the resonance, in a region where n (ω ≈ ω 0 ) is small and the phase velocity,
c/n (ω 0 ) is greater than the speed of light. The normalized frequency is ω/ω T .
v/c
0.25
0
1.4 1.55 1.7 1.85 2
Normalized frequency
The plasma presents a different aspect of the wave propagation problem. The group velocity can exceed the speed of
light in the region where the index of refraction (see Fig. 7) is small and slowly varying. However, in this region there is
no wave propagation as illustrated by Fig.10 . This shows the phase change for the wave as the intensity falls by 1/e. The
normalized frequency is ωτ , where ω p τ = 10.
3
Phase change, radians
0
0 2 4 6 8 10 12
normalized frequency
50
Section 2.4 Propagation of plane waves in a medium
For reference the decay length, in units of cτ , for the transverse wave in a plasma is shown in Fig. 11 . For metallic
plasmas ω p ≈ 1016 rad/s and the figure would correspond to a system with a damping constant of τ ≈ 10−15 s.In this case
cτ ≈ 100 nm.
0
0 5 10 15 20
Normalized frequency
2.4.2 Causality Our concept of causality sequences events using the speed of light as the limiting factor. Let the time
and location of two events be (r, t) and (r0 , t0 ) . The event at (r0 , t0 ) is in the past of the event at (r, t) if |r − r0 | < c (t − t0 ) .
This means that t is so much larger than t0 that the signal (emitted at (r0 , t0 )) has long since reached the observation point r.
If t is not quite large enough (but t > t0 ) the event is in the present of (r, t) and |r − r0 | > c |t − t0 |. That is, the signal has
not yet reached the observation point, r. The event is in the future of (r, t) if t0 > t. and |r − r0 | < c (t0 − t). This
means that the signal has not yet been generated.. Causality requires that a given event can only be inuenced by events
occurring in its past. Therefore no signal may propagate faster than the speed of light. In particular, a wave in a dispersive
medium will propagate with a speed which does not exceed the vacuum speed of light. This places constraints on the index
of refraction and the electric permittivity. Let u (x, t) be a wave propagating in a medium with a permittivity ε (ω) and let
1/2
n (ω) = ε (ω) . For example
Z · µ ¶ µ ¶¸
ωn (ω) x ωn (ω) x
u (x, t) = e−iωt A (ω) exp i + B (ω) exp −i dω (2.197)
c c
and similarly
Z ∞
∂ ωn (ω) ωn (ω)
u (x, t) |x0 =o = e−iωt [i A (ω) − i B (ω)]dω (2.199)
∂x0 −∞ c c
Z ∞ Z ∞ Z ∞
iω 0 t ∂ ωn (ω) −i(ω−ω0 )t
e u (x, t) |x0 =o dt = e i [A (ω) − B (ω)]dωdt
−∞ ∂x0 −∞ −∞ c
Z ∞
ωn (ω)
= 2π [A (ω) − B (ω)]i δ(ω − ω 0 )dω and
−∞ c
Z ∞
c 0 ∂
−i 0 eiω t u (x, t) |x0 =o dt = [A (ω 0 ) − B (ω 0 )]
ω n (ω 0 ) 2π −∞ ∂x0
Finally,
Z ∞
1 0 c ∂
A (ω 0 ) = eiω t {u (0, t) − i 0 u (x, t) |x0 =o }dt (2.200)
2 · 2π −∞ ω n (ω 0 ) ∂x0
Z ∞
1 0 c ∂
B (ω 0 ) = eiω t {u (0, t) + i 0 u (x, t) |x0 =o }dt (2.201)
2 · 2π −∞ ω n (ω 0 ) ∂x0
In the integrals for A (ω) and B (ω) let ω 0 be replaced by ω. And let t be replaced by t0 to distinguish it from the value of
time on the left hand side in the following expression. :
Z ∞ Z ∞ µ ¶½ ¾
1 ωn (ω) x
−iω (t−t0 ) 0 ic
u (x, t) = e exp i u (0, t ) − [∂u (x , t ) /∂x ]x0 =0 dωdt0
0 0 0
4π
−∞ −∞ c ωn (ω)
Z ∞Z ∞ µ ¶½ ¾
1 −iω (t−t0 ) ωn (ω) x ic
+ e exp −i 0
u (0, t ) + [∂u (x , t ) /∂x ]x0 =0 dωdt(2.202)
0 0 0 0
4π −∞ −∞ c ωn (ω)
Next, the integration over ω can be converted into a contour integration in the complex ω plane.
Z Z µ ¶½ ¾
1 0 ωn (ω) x ic
u (x, t) = e−iω(t−t ) exp i u (0, t0 ) − [∂u (x0 , t0 ) /∂x0 ]x0 =0 dωdt0
4π C c ωn (ω)
Z Z µ ¶½ ¾
1 0
−iω (t−t ) ωn (ω) x ic
+ e exp −i 0
u (0, t ) + [∂u (x , t ) /∂x ]x0 =0 dωdt0 (2.203)
0 0 0
4π C c ωn (ω)
with C denoting the contour which passes above any poles or cuts on the real ω axis. To observe the constraints imposed by
causality it suffices to consider the following expressions for x > 0,
Z µ ¶
1 −iω (t−t0 ) ωn (ω) x ic
0
g (x, t − t ) = − e exp i dω (2.204)
4π C c ωn (ω)
and
Z µ ¶
∂ 1 0 ωn (ω) x
g (x, t − t0 ) = f (x, t − t0 ) = e−iω(t−t ) exp i dω (2.205)
∂x 4π C c
A general property of ε (ω) and n (ω) is that for large |ω| they approach one. To obtain well behaved integrands for
large |ω| we will subtract off this asymptotic dependence. But first we consider the functions with n (ω) = 1,call them
g0 (x, t − t0 ) and f0 (x, t − t0 ) . The following are evaluated using a contour in the lower half plane and the Cauchy residue
52
Section 2.4 Propagation of plane waves in a medium
where the path of integration is taken to be just below the real ω axis. To use Cauchy’s theorem to evaluate the integral for
t < 0 we would attempt to close the path in the upper half plane. The difficulty is that ω −1/2 ‘cuts the plane’. Let ω = |ω| eiφ
1/2
with −π < φ ≤ π then as ω approaches the negative real axis from below ω 1/2 → −i |ω| while as ω approaches the
53
Section 2.4 Propagation of plane waves in a medium
1/2
negative real axis form above ω 1/2 → +i |ω| . This discontinuity occurs because the function z 2 maps the complex plane
with −π < φ ≤ π into the ‘two sheets’ −2π < φ0 ≤ 2π.
To avoid the ‘second sheet’ when the path is closed, the path must not cross the cut. When the path is closed in the upper
half plane the segments of the path are, in the limit R → ∞, (1) −R − iδ → +R − iδ, (2) z = R · eiφ from +R − iδ to
−R + iδ, (3) −R + iδ → iδ, (4) z = δ · eiφ from +iδ to −iδ, and (5) −iδ → −R − iδ. This path is shown in Fig. . We will
now evaluate C (t) . For t > 0 C (t) = 0 and for t < 0 we obtain the integral of the discontinuity if the integrand across the
cut.
Z 0 Z ∞
r
exp (−iωt) 2i 2π
C (t) = − 1/2
(−2i) dω = √ u−1/2 exp (−iu) du = (1 + i) , t<0 (2.213)
−∞ (−ω) −t 0 −t
We note that a cut in the upper half plane would lead to a nonzero value for the integral in the case that the path was to be
closed in the upper half plane.
.................................................................................................................
.................................................................................................................
Omit: Assignment 8: The claim is that the cut in the plane defining ω 1/2 can be taken along any line going from zero
to infinity in the upper half plane. (Note the physics of a problem determines whether the path of integration is taken above
or below the cut. There’s no physics in this problem.) Check whether this claim is correct by taking ω = |ω| exp (iφ) with
−3π/2 < φ ≤ π/2.
.................................................................................................................
.................................................................................................................
1/2
The requirement that γ (ω) have no cuts or poles in the upper half complex ω plane is satisfied if n (ω) = ε (ω) has no
poles or cuts in the upper half complex plane. In this case ε (ω) can neither have a pole nor a zero in the upper half plane.
This condition is satisfied by the model permittivities for the ionic crystals and for the charged plasmas.
A similar argument applies to f (x, t − t0 ) − f0 (x, t − t0 ) . This function is the fourier transform
Z · µ ¶ ¸
1 0 ω [n (ω) − 1] x/c
f (x, t − t0 ) − f0 (x, t − t0 ) = e−iω(t−t −x/c) exp i − 1 dω (2.214)
4π C c
Again the function being transformed vanishes as ω → ±∞ and the integral will be zero if t > 0 and ε (ω) has no poles or
zeros in the upper half plane.
54
Section 2.4 Propagation of plane waves in a medium
2.4.3 Wave dispersion in a medium In order to examine the dispersion character for a wave in a medium we
take u (0, t) = u0 δ (t) and [∂u (x, t) /∂t]x=0 = 0. These boundary conditions will provide identical waves travelling in the
+x and −x directions. Since the Fourier transform of a delta function is a constant this acts as a ‘white light’ source. From
Eq. 200,
Z Z µ ¶ µ ¶
1 0 ωn (ω) x ωn (ω) x
u (x, t) = e−iω(t−t ) [exp i + exp −i ]u (0, t0 ) dωdt0 (2.215)
4π C c c
Z Z µ ¶
1 −iω (t−t0 ) ωn (ω) x
u (x, t) = e cos u0 δ (t0 ) dωdt0 , t > 0
2π C c
Z µ ¶
u0 ωn (ω) x
= e −iωt
cos dω, t > 0 (2.216)
2π C c
Since u (x, t) = u (−x, t) it suffices to consider the wave travelling in either the +x or −x direction.
Z µ · ¸¶ Z µ · ¸¶
u0 n (ω) x u0 n (ω) x
u (x, t) = exp −iω t − dω + exp −iω t + dω t > 0 (2.217)
4π C c 4π C c
= u+ (x, t) + u− (x, t)
We will chose the wave travelling in the +x direction for which the transform is
Z µ · ¸¶
u0 n (ω) x
u+ (x, t) = exp −iω t − dω, t > 0 (2.218)
4π C c
Z ∞ µ · ¸¶
u0 n (ω) x
= exp −iω t − dω
4π −∞ c
Z ∞ µ · ¸¶ µ · ¸¶
u0 n (ω) x n (ω) x
= [cos −ω t − + i sin −ω t − ]dω
4π −∞ c c
Since n (ω) = n (−ω) only the real part of the last expression survives to give
Z ∞ µ · ¸¶
u0 n (ω) x
u+ (x, t) = Re exp −iω t − dω, t > 0 (2.219)
2π 0 c
In the following we will assume, incorrectly, that the imaginary part of the permittivity vanishes. That is,
ω ω
k = [Re ε(ω) + i0]1/2 = [n (ω) + ikap(ω)] (2.220)
c c
so that
1/2 £ 2 ¤1/2
ωn (ω) = εi |ω − ω 2p | (2.222)
55
Section 2.4 Propagation of plane waves in a medium
This will allow us to use a special case of the ‘method of steepest descent’4 which is less tedious. The method is known as
the ‘stationary phase approximation’. The approximation will be illustrated using the electric permittivity of an undamped
plasma, τ → ∞. q
For an undamped plasma with a plasma frequency ω p we have that ωn (ω) = ω 2 − ω 2p for ω > ω p . For −ω p ≤ ω ≤ ω p
q
we obtain ωn (ω) = i ω 2p − ω 2 . It is convenient to define a time t0 = x/c which is the time it takes for a light signal
travelling in a vacuum to go from the origin to the observation point x. We also let ω = ξω p and t = υt0 . We have
determined that this system satisfies causality so that υ ≥ 1. With these definitions the function can be written as
Z ∞ µ · q ¸¶
u0 ω p 2
u+ (x, t) = Re exp −iω p t0 υξ − ξ − 1 dξ
2π 1
Z 1 µ · q ¸¶
u0 ω p
+ Re exp −ω p t0 iυξ + 1 − ξ 2 dξ (2.223)
2π 0
Our problem is to obtain an estimate for the value of this function if ω p t0 À 1. The second integral is exponentially decaying
with distance and so will be assumed to be negligible.
The phase, φ(ω), occurring in the first integral is ω p t0 f (υ, ξ) with
q
f (υ, ξ) = υξ − ξ 2 − 1 (2.224)
Fig. 13 gives f (υ, ξ) = φ(ω)/ω p t0 for three observation times, 1.1t0 , 1.2t0 , and 1.3t0 .
φ /(ω pt 0)
3.50
3.00
t = 1.3 t 0
2.50
2.00
1.50 t = 1.2 t 0
1.00
0.50 t = 1.1 t 0
0.00
1 3 5 7 9
ω/ω p
.Fig. 13
Regions in which the phase varies ‘rapidly’ will contribute little to the integral. For large ω p t0 the major contribution to
for large α. The saddle points of u (z) , z complex, are located and the path of integration is distorted so as to pass through these saddle points. In going
through a saddle point the path is along the line for which the curvature is negative. The integral is approximated by using only the neighborhoods of the
saddle points and approximating u (z) by a quadratic in the distance from the saddle point. The classic reference for this is Courant & Hilbert, Methods of
Mathematical Physics, Vol. 1 (Interscience Publishers, pp526532) Another reference is Approximation methods in quantum mechanics by A. B. Migdal
and V. Krainov (pp 68, W. A. Benjamin, 1969)
56
Section 2.4 Propagation of plane waves in a medium
the integral will come from the region where the magnitude of the phase is minimum or has zero slope. This occurs at
∂f ξ
=0=υ− p 2 at ξ = ξ o (2.225)
∂ξ ξ −1
ω υ 1
ξo ≡ =√ .= r (2.226)
ωp 2
υ −1 to
1 − [ ]2
t
1
ξ 2o − 1 = (2.227)
υ2 −1
10
t/t 0
.Fig. 14
For t → t0 the frequency of the minimum of the magnitude of the phase approaches infinity. On the other extreme as
t → ∞ the frequency of the minimum approaches the plasma frequency. We deduce then that the frequency observed at the
observation point starts high and decreases with time. The range of frequencies which contribute to the observed signal also
decreases with time. This can be obtained by considering the reciprocal of the curvature of the phase as a function of ω.
∂ 2 φ (ω; x, t) t0 2
= (ξ − 1)−1.5
∂ω 2 ωp
t0 ¡ 2 ¢3/2
= ν −1 at ξ = ξ o
ωp
∂2φ ¡ ¢3/2
= ω p t0 ν 2 − 1 at ξ = ξ o (2.228)
∂ξ 2
57
Section 2.4 Propagation of plane waves in a medium
4
Reciprocal phase
curvature at minimum
3
t/t 0
.Fig. 15
This is shown in Fig. 15 . The approximation to u+ (x, t) is obtained by expanding the phase about the minimum,
keeping only the quadratic terms5
1
φ(ξ) = φ(ξ o ) + φ0 (ξ o )(ξ − ξ o ) + φ00 (ξ o )(ξ − ξ o )2 + ...
2
1
≈ φ(ξ o ) + φ00 (ξ o )(ξ − ξ o )2
2
p 1
= ω p t0 υ2 − 1 + ω p t0 [υ 2 − 1]3/2 (ξ − ξ o )2 (2.230)
2
are Fresnel integrals. (Abramowitz and Stegun, Hanbook of Mathematical Functions, p.300, Dover) Both integrals go to 0.5 as z → ∞.
58
Section 2.4 Propagation of plane waves in a medium
Z ∞ · ¸
u0 ω p 1 00 2
u+ (x, t) ≈ Re [ exp (−iφ(ξ o )) exp −i φ (ξ o ) (ξ − ξ 0 ) dξ ]
2π 1 2
Z ∞ h π i
u0 ω p 1 00 −1/2
= Re [ exp (−iφ(ξ o )) [ φ (ξ o )] exp −i W 2 dW ]
2π π W (ξ=1) 2
Z ∞
u0 ω p 1 π π
= Re [ exp (−iφ(ξ o )) [ φ00 (ξ o )]−1/2 {cos( W 2 ) − i sin( W 2 )}dW ] (2.232)
2π π W (ξ=1) 2 2
where we used
1 00
W2 = φ (ξ o ) (ξ − ξ 0 )2 and (2.233)
π
1
W (ξ = 1) = [ ω p t0 ]1/2 [υ 2 − 1]3/4 (1 − ξ 0 )2
π
1 t2 2
= [ ω p t0 ]1/2 [ 2 − 1]3/4 (1 − ξ 0 )
π to
t
≈ 0 for large where ξ 0 ≈ 1 (2.234)
to
u0 ω p 1
u+ (x, t) ≈ Re [ exp (−iφ(ξ o )) [ φ00 (ξ o )]−1/2 (1 − i)/2] (2.235)
2π π
" ¡ √ ¢ #
u0 exp −iω p t0 υ 2 − 1
u+ (x, t) ≈ √ Re (1 − i)
4 πt0 (υ 2 − 1)3/4
√ " ¡ √ ¢ #
2u0 exp −iω p t0 υ 2 − 1
= √ Re exp(iπ/4)
4 πt0 (υ 2 − 1)3/4
µ q ¶
2 2
cos ω p t − (x/c) + π/4
u0
= √ ³ ´3/4 (2.236)
2 2π (x/c)1/4 t2 − (x/c)
2
This approximation is good for ω p t0 À 1. Considering the range of plasma frequencies, from 109 rad/s for gaseous plasmas
through 1016 rad/s for metallic plasmas, the distance of the observation point from the source should be much greater than
10 cm for the gaseous plasmas and 1 µm for the metals. The plot in Fig. 16 shows u+ (x, t) for a metallic plasma with
ω p = 1016 rad/s and ω p to = 1000. at three times, t. Note that the frequency decreases as time increases.
59
Section 2.5 Analyticity and the KramersKronig relationship
6
4.834
u( x, to )
u( x, 2⋅ to ) 0
u( x, 4⋅ to )
− 5.761 6
0 5 10 15 20 25 30
0 x 30
−6
10
x in microns
(2.237)
Figure 16
2.5 Analyticity and the KramersKronig relationship We have determined that causality requires ε (ω)
be analytic and have no zeros in the upper half complex ω plane. In the case that the local approximation holds, this can be
seen by considering the causal expression
Z ∞
D (r, t) = E (r, t) + G (τ ) E (r, t − τ ) dτ (2.238)
0
In this expression G (τ ) is a property of the system and is generally expected to vanish as τ → ∞. (The exception to this
behavior is furnished by the electrical conductors for which G (τ ) → 4πσ 0 , the DC conductivity6 .) The Fourier transform,
in time, of this expression along with the relationship between the frequency components of D and E yields
Z ∞ Z ∞ Z ∞Z ∞
D (r, t) eiωt dt = E (r, t) eiωt dt + G (τ ) E (r, t − τ ) eiωt dtdτ (2.239)
−∞ −∞ −∞ 0
Z ∞Z ∞
D (r, ω) = E (r, ω) + G (τ ) eiωτ E (r, t − τ ) eiω(t−τ ) d(t − τ )dτ
−∞ 0
Z ∞
ε (ω) E (r, ω) = E (r, ω) + G (τ ) eiωτ E (r, ω) dτ
0
Z ∞
ε (ω) − 1 = G (τ ) exp (iωτ ) dτ (2.240)
0
For τ → 0 G (τ ) → 0 which can be interpreted as demonstrating the finite time it takes for a system to repond to a stimulus. While for τ → ∞
G (τ ) → 4πσ 0 .
60
Section 2.5 Analyticity and the KramersKronig relationship
Since D and E are real G (τ ) will also be real. It follows, as previously noted, that
∗
ε (ω) = ε (−ω) . (2.241)
It is also possible to obtain a relationship between G and its derivatives at τ = 0 and the high frequency behavior of
ε (ω) . To arrive at the relationship we note that
Z ∞
G (τ ) exp (iωτ ) dτ
0
µ Z ∞ ¶
1 dG (τ )
= G (τ ) exp (iωτ ) |∞
0+ − exp (iωτ ) dτ (2.242)
iω 0 dτ
and
Z ∞
dn G (τ )
exp (iωτ ) dτ
0 dτ n
µ n Z ∞ n+1 ¶
1 d G (τ ) d G (τ )
= ∞
exp (iωτ ) |0+ − exp (iωτ ) dτ (2.243)
iω dτ n 0 dτ n+1
dn+1 G(τ )
For all materials which are not DC conductors (and for which limn→∞ dτ n+1 = 0)
X∞ n
(−1) dn−1 G (τ 0 )
ε (ω) − 1 = | 0 + (2.244)
n=1
(iω)n dτ 0 n−1 τ =0
The asymptotic behavior of ε (ω) for large ω can be deduced from this expression. It is reasonable to expect that the
response of the system takes a finite time, i.e., the system does not instantaneously change. In the case that this is true we
have that G (0) = 0 and the first two terms in the expansion are
ε (ω) − 1 ≈ −ω −2 G0 (0) − iω −3 G00 (0) (2.245)
−1 0
Re(ε (ω) − 1) ≈ G (0) (2.246)
ω2
−1 00
Im(ε (ω) − 1) ≈ G (0) (2.247)
ω3
That is, the real part of ε (ω) − 1 vanishes as ω −2 and the imaginary part of ε (ω) vanishes as ω −3 .
Finally, for complex ω in the upper half (Im ω > 0)
Z ∞ Z ∞ Z ∞
ε (ω) − 1 = iωτ
G (τ ) e dτ ≤ G (τ ) |eiτ Re ω −τ Im ω
|e dτ ≤ G (τ ) dτ (2.248)
0 0 0
R∞
If 0 G (τ ) dτ is finite then ε (ω) − 1 will be finite and an analytic function of complex ω in the upper half (Im ω > 0)
complex plane. This has some important consequences. The ε (ω) − 1 for ω real exists and is the limit of a function which
is analytic in the upper half plane.
2.5.1 The KramersKronig relations Since ε (ω) − 1 is an analytic function in the upper half complex plane
then, by Cauchy’s residue theorem,
I
1 ε (z 0 ) − 1 0
ε (z) = 1 + dz (2.249)
2πi C z0 − z
with the imaginary part of all points on the curve C restricted to be greater than or equal to zero. Let the path lie along the
real axis and let it be closed by the semicircle of infinite radius. (ε (ω) − 1 will vanish on this semicircle.) In addition let
z = ω + iδ,with 1 À δ > 0 then
Z ∞
1 ε (ω 0 ) − 1
ε (ω) = 1 + lim dω 0 (2.250)
δ→0 2πi −∞ ω 0 − ω − iδ
61
Section 2.5 Analyticity and the KramersKronig relationship
This can be rewritten using the relationships between path integrals, principal value integrals, and delta functions. The
principal value of the integral, where f (z 0 ) is analytic in upper half plane, is:
Z I Z
f (z 0 ) 0 f (z 0 ) 0 f (z 0 ) 0
P dz = dz − dz (2.251)
z0 − z C z0 − z Cδ z0 − z
where Cδ is a circular curve of radius δ 0 which ”surrounds” the point, z. If z lies on the real axis then Cδ is a semicircle of
radius δ 0 taken below the point z and
Z I Z
f (z 0 ) 0 f (z 0 ) f (z 0 ) 0
P dz = 0
dz − dz (2.252)
z0 − z 0
C z − (z + iδ) Cδ z − z
0
Z
f (z 0 )
= 2πif (z + iδ) − 0
d(z 0 − z) (2.253)
Cδ z − z
Z 2π ¡ ¢
f z + δ 0 eiφ
= 2πif (z + iδ) − d(δ 0 eiφ ) with z 0 − z = δ 0 eiφ (2.254)
π δ 0 eiφ
0 iφ
Z 2π f (z) + f 0 (z) δ e + ...
= 2πif (z + iδ) − 2 idφ (2.255)
π 1
δ0
= 2πif (z + iδ) − if (z)π − f 0 (z) [ei2π − eiπ ] + ... (2.256)
2
Applying this to our problem, the integral taken below the pole equals the principal part of the integral plus πi times the
integral with a delta function
Z
1 ε (ω 0 ) − 1 0
ε (ω) = 1 + [P dω + iπ(ε (ω) − 1)]
2πi ω0 − ω
Z Z
1 ε (ω 0 ) − 1 0 ε (ω 0 ) − 1
= 1+ [P dω + πiδ(ω 0 − ω)dω 0 ]
2πi ω0 − ω ω0 − ω
Z
1 ε (ω 0 ) − 1 0 1
= 1+ P dω + [ε (ω) − 1]
2πi ω0 − ω 2
or
Z
1 ε (ω 0 ) − 1 0
ε (ω) = 1 + P dω (2.258)
πi ω0 − ω
Using the fact that the integrand vanishes on the upper half circle at ω 0 = ∞, this yields relationships between the real and
imaginary parts of ε (ω)
Z ∞
1 Im ε (ω 0 ) 0
Re ε (ω) = 1 + P 0
dω (2.259)
π −∞ ω − ω
Z ∞
1 Re [ε (ω 0 ) − 1] 0
Im ε (ω) = − P dω (2.260)
π −∞ ω0 − ω
62
Section 2.5 Analyticity and the KramersKronig relationship
this case, as seen in the lectures covering the permittivity of a charged plasma, there is a singularity at ω = 0. The singularity
63
Section 2.5 Analyticity and the KramersKronig relationship
It follows that
Z ∞
2 1 1
Re ε (ω) − Re ε (ω 1 ) = P ω 0 Im ε (ω 0 ) [ 0 2 − 02 ]dω 0
π 0 ω − ω 2 ω − ω 21
Z ∞
2 ¡ 2 ¢ ω 0 Im ε (ω 0 )
= 2
ω − ω1 P dω 0 (2.267)
π 0 (ω 0 2 − ω 2 ) (ω 0 2 − ω 21 )
This form for ε (ω) reduces the integral’s dependence on the values of Im ε (ω) at large ω. This is known as a ‘subtracted
dispersion relation’. This process of subtraction can be repeated for each known value of ε (ω) . A similar relationship can
be generated, starting with Eq. 2.262, for the imaginary part ε (ω) .
2.5.2 Sum rules The KramerKronig relations can be used to obtain some general properties of the integrals¡ of the ¢
real and imaginary parts of the permittivity. We have noted that, at high frequency, the permittivity varies as ω −2 + O ω −3 .
Physically at high frequencies the binding forces for the electrons becomes negligible and the system responds as a plasma.
With this interpretation we define the plasma frequency for the system
£ ¤
ω 2p ≡ lim ω 2 (1 − ε (ω)) (2.268)
ω→∞
¡ ¢
If Im ε (ω) varies as ω −3 + O ω −4 for high frequencies then Eq. 2.261 gives
£ ¤
ω 2p = lim ω 2 (1 − Re ε (ω)) − ω 2 i Im ε (ω) (2.269)
ω→∞
£ ¡ ¢¤
= lim ω 2 (1 − Re ε (ω)) − ω 2 i(ω −3 + O ω −4 )
ω→∞
· Z ∞ 0 ¸
2 ω Im ε (ω 0 ) 0 ¡ −2 ¢
= lim −ω 2 P dω − i(ω −1
+ O ω )
ω→∞ π 0 ω0 2 − ω2
Z ∞ 0
2 0
ω Im ε (ω ) 0 ¡ ¢
= lim − P 02 dω − i(ω −1 + O ω −2 )
ω→∞ π 0 ω
[ 2 − 1]
ω
Z ∞
2
ω 2p = ω Im ε (ω) dω (2.270)
π 0
It is interesting to consider the relationship between the coefficients in Sellmeier’s equation, Eq. 2.122 , for the
64
Section 2.5 Analyticity and the KramersKronig relationship
permittivity and the plasma frequency defined in Eq. 2.270.The sum rule requires that
Z
2X ∞ Am
2
ωp = ω Im 2 dω (2.271)
π m 0 ω m − iγ m ω − ω 2
Z
2X ∞ Am [ω 2m − ω 2 + iγ m ω]
= ω Im[ 2 dω
π m 0 [ω m − ω 2 − iγ m ω][ω 2m − ω 2 + iγ m ω]
Z
2X ∞ Am γ m ω
= ω[ 2 dω
π m 0 (ω m − ω 2 )2 + (γ m ω)2
Z
1X ∞ Am γ m ω 2
= dω (2.272)
π m −∞ (ω 2m − ω 2 )2 + (γ m ω)2
and evaluate it using Cauchy’s residue theorem. Closing the path in the upper half complex ω plane we obtain the residues of
the poles in the upper half plane (at ω + and ω − )
X A γ ω2
ω 2p = 2i ¡ m m ∗+¢ ¡ ¢
m
(ω + − ω − ) ω + − ω + ω + − ω ∗−
X A γ ω2
+2i ¡ m m ∗−¢ ¡ ¢ (2.274)
m
(ω − − ω + ) ω − − ω + ω − − ω ∗−
X Am γ m ω 2+ Am γ m ω 2−
= 2i [ γm ¡ ¢ + ¡ ¢ γ
m (ω + − ω − ) 2i ω + − ω ∗− (ω − − ω + ) ω − − ω ∗+ 2i m
" 2 # 2
X Am ω 2+ Am ω 2−
= 2 ¡ ¢+ ¡ ¢
m
(ω + − ω − ) ω + − ω ∗− (ω − − ω + ) ω − − ω ∗+
X · Am ω + Am ω −
¸
= − since ω + = −ω ∗− and ω ∗+ = −ω −
m
(ω + − ω − ) (ω + − ω − )
X
= Am (2.275)
m
The definition of the plasma frequency given in Eq. 2.268 is quite general and would reect the contributions of the
responses of all charged particles in the system. The coefficients in Sellmeier’s equation, which approximates the permittivity
as a sum of resonances, gives the ‘strength’ of each resonance. In principle the sources of the resonances include optically
active phonons, plasmons, electronic excitations, electronic ionizations, etc.
¡ ¢ ¡ ¢
A second sum rule is obtained if Re ε (ω) − 1 ∼ −ω 2p /ω 2 + O ω −4 and Im ε (ω) ∼ O ω −3 for large ω. The
asymptotic relationship between the real and imaginary parts of ε (ω) can be obtained using Equation 2.262. We write this
expression as
"Z Z ∞ #
ωm
2 Re [ε (ω 0 ) − 1] 0 Re [ε (ω 0 ) − 1] 0
Im ε (ω) = 2 dω + P 2 dω (2.276)
πω 0 1 − (ω 0 /ω) 0
ω m 1 − (ω /ω)
65
Section 2.5 Analyticity and the KramersKronig relationship
with ω À ω m and ω m sufficiently large that the asymptotic form for the real part of ε (ω) provides a valid approximation.
The approximated equation is
Z ωm h i
2 2
Im ε (ω) ≈ Re [ε (ω 0 ) − 1] 1 + (ω 0 /ω) dω 0
πω 0
Z ∞
2 1 £ 2 02 ¡ ¢¤
+ P 2 −ω p /ω + O ω 0−4 dω 0
πω ωm 1 − (ω /ω)
0
·Z ωm ¸
2 ¡ ¢
Im ε (ω) ≈ Re [ε (ω 0 ) − 1] dω 0 − [ω 2p /ω m + O ω −2 (2.277)
πω 0
Note that for an integral over finite limits with f (x) = f (−x)
Z "Z Z #
a −δ a
f (x) f (x) f (x)
P dx = lim dx + dx
−a x δ→0 −a x δ x
· Z a Z a ¸
f (−x) f (x)
= lim − dx + dx = 0 (2.278)
δ→0 δ x δ x
¡ ¢
Since Im ε (ω) ∼ O ω −3 for large ω it follows that
Z ωm
Re [ε (ω 0 ) − 1] dω 0 − ω 2p /ω m = 0 (2.279)
0
or
Z ωm
ω −1
m Re ε (ω 0 ) dω 0 = 1 + ω 2p /ω 2m (2.280)
0
This is sometimes called a superconvergence relation. This latter relation, Eq. 2.280, does not hold for conductors while the
sum rule for oscillator strengths, Eq. 2.270, will hold for conductors.
2.5.3 Some useful relations between path integrals The KramersKronig relations involve principal value
integrals. For these integrals the integrand will have simple poles on the real axis and the integral is taken along the real axis
excluding the location of the pole.
66
Section 2.5 Analyticity and the KramersKronig relationship
.Fig. 16
For example let f (x) have no singularities on the real axis then
Z "Z Z #
∞ −δ ∞
f (x) f (x) f (x)
P dx = lim dx + dx (2.281)
−∞ x δ→0 −∞ x δ x
Often it is easier to evaluate the integral using Cauchy’s residue theorem. This requires a closed path and the principal value
integration leaves a gap in the path. This problem is surmounted using the identity illustrated in Fig. 16 . The integrand is
assumed to have a simple pole located at the ‘x’ on the real axis The identity holds because one can show that the integrals
along the half circles in the two contours cancel. In this case
Z "Z Z #
∞ ∞ ∞
f (x) 1 f (x) f (x)
P dx = dx + dx (2.282)
−∞ x 2 −∞, C+ x −∞, C− x
The principal value integral has been replaced by two integrals but Cauchy’s residue theorem can often be used to evaluate
each integral.
67
Section 2.6 Vector characteristic of electromagnetic wave
.Fig.17
Another relationship between the path integrals is often useful. This identity is illustrated in Fig. 17 where again ‘x’
marks the location of the pole. The direction of integration along the circular path around the pole has been selected so as to
cancel the half circle for C+ and add the half circle for C− .
If both identities are combined the principal value integral is seen to equal (1) the integral along path C+ plus one half
the integral around the pole or (2) the integral along the path C− minus one half the integral around the pole.
2.6 Vector characteristic of electromagnetic wave The final property of electromagnetic waves we
shall consider is their vector characteristic. In an isotropic medium the electric field is perpendicular to the direction of
propagation of the wave. A plane wave with the electric field given by E(r,t) = E1 (r, t) ε1 is said to be linearly polarized
with polarization vector ε1 . Similarly E(r,t) = E2 (r, t) ε2 (with ε1 · ε2 = 0) describes a linearly polarized wave with
polarization vector ε2 . A general plane electromagnetic wave with wave vectork (k · εi = 0, k×ε1 = kε2 ) is given by
ε± · ε∗± = 1
ε± · ε∗∓ = 0
ε∗± = ε∓
68
Section 2.6 Vector characteristic of electromagnetic wave
A wave described by E(r, t) = E0 ε± exp (i [k · r − ωt]) is a circularly polarized wave. The components of this wave are
With the upper sign of ε± the wave is said to be left hand circularly polarized and with the lower sign it is right hand
circularly polarized. This is determined by the direction of rotation of the electric field in the advancing wave (at a fixed time
not at a point). An observer (when facing the oncoming wave) sees the E rotating counterclockwise ( clockwise) at a fixed
space point for the ε+ (ε− ) polarization In modern terminology the left hand circularly polarized wave has positive helicity
while the right hand circularly polarized wave has negative helicity. The sign is the sign of the projection of the angular
momentum of the wave on the direction of wave propagation.
The general polarized wave has elliptical polarization. It is most easily seen using the circularly polarized waves. The
semimajor axis is reached when the countercirculating electric fields are aligned. Its length is therefore
√
a = 2 (|E+ | + |E− |)
. The semiminor axis is reached when the fields are antialigned resulting in the length
√
b = 2 ||E+ | − |E− ||
. If the waves are in phase the major axis is along ε1 and the minor axis is along ε2 . Let φ+ be the phase of the positive
helicity wave and φ− the phase of the negative helicity wave. If θ± are the angles between the electric fields and ε1 then
θ+ = k · r − ωt + φ+ and
θ− = −k · r + ωt − φ− .
The alignment of these two vectors resulting in the semimajor axis occurs for θ+ − θ− = 0, 2N π. This has solutions
¡ ¢ ¡ ¢
k · r − ωt = −0.5 φ+ + φ− , −0.5 φ+ + φ− + N π
2.6.1 Stoke’s parameters The polarization of an electromagnetic wave can be determined by intensity
measurements. These measurements yield the magnitude of the amplitudes of ε± and ε1,2 and the relative phases of these
amplitudes. In terms of the field expression the amplitudes are
These parameters are the Stoke’s parameters. They are not all independent since they only depend on the relative phase
of the amplitudes. Ideally they satisfy
s20 = s21 + s22 + s23 (2.295)
however there is no such thing as a monochromatic wave. Any spread in the frequencies will lead to a variation in the
amplitudes, both the magnitudes and phases. Time average measurements yield that
s1 = s2 = s3 = 0 (2.297)
The Stokes parameters are quadratic in the field strength and can be determined through intensity measurements only, in
conjuction with a linear polarizer and a quarterwave plate or the equivalent. Their measurement determines completely the
state of polarization of the wave.
.................................................................................................................
.................................................................................................................
Assignment 10: Jackson 7.1
.................................................................................................................
.................................................................................................................
2.7 Reection and refraction at a plane interface between dielectrics Consider a plane
electromagnetic wave (wave vector k and angular frequency ω) travelling in a medium with a permittivity εi (ω) and
permeability µi (ω) . We assume that this medium fills the space z < 0 and that the region z > 0 is filled with a medium
with a permittivity εt (ω) and permeability µt (ω) . From our investigations of Maxwell’s equations we have learned that
the tangential components of E and H and the normal components of D and B are continuous at the interface between the
materials. Without any loss of generality we can assume that the electric field part of the wave is described by the real part
of
Ei (r, t) = Ei exp [i (k1 x + k3 z − ωt)] , (2.298)
i.e., the plane of incidence is perpendicular to the y axis. The continuity conditions at the interface require that the transmitted
and reected electric fields have the same spatial variation as the incident field along the planar interface,
with
ω2
k12 + k32 = εi (ω) µi (ω) (2.301)
c2
ω2
k12 + k30 2 = εt (ω) µt (ω) 2 (2.302)
c
This requirement is seen to yield Snell’s law of refraction along with the ‘law of reection’ at interfaces between media.
70
Section 2.7 Reection and refraction at a plane interface between dielectrics
To relate the various fields to the electric fields we consider the spatial and temporal fourier transform of the source free
Maxwell’s equations along with the constituent equations for an infinite medium.
c
B0 = k × E0 (2.303)
ω
H0 = µ (ω)−1 B0 (2.304)
D0 = ε (ω) E0 (2.305)
It is convenient to consider two cases, the first with the incident electric field perpendicular to the plane of incidence
Ei = Ei e2 (2.306)
and the second with the incident electric field parallel to the plane of incidence
k3 e1 − k1 e3
Ei = Ei (2.307)
k
In each case the electric fields of the transmitted and reected waves will lie in the same planes as the incident wave.
2.7.1 Electric field perpendicular to plane of incidence With the incident electric field perpendicular to
the plane of incidence the electric fields of the transmitted and reected waves will have the amplitudes
Et = Et e2 (2.308)
Er = Er e2 (2.309)
The continuity requirement for the component of the electric field which is tangential to the interface gives
Ei + Er = Et (2.310)
The tangential components of the magnetic field vector, H, for the incident, transmitted8 , and reected waves are
The continuity requirement for the component of the magnetic field which is tangential to the interface gives
−1 −1
µi (ω) k3 [Ei − Er ] = µt (ω) k30 Et (2.314)
71
Section 2.7 Reection and refraction at a plane interface between dielectrics
Generally the ratio of this reected component to the incident component gives the fraction of the incident intensity which is
reected
¯ ¯
¯ µ (ω) k3 − µi (ω) k30 ¯2
R⊥ = ¯¯ t ¯ (2.319)
µt (ω) k3 + µi (ω) k30 ¯
The component of the average transmitted Poynting vector which is perpendicular to the interface is
µ 0 ¶¯ ¯2
k3 ¯ 2µt (ω) k3 ¯
[St ]avg = Re ¯ ¯ |Ei |2 (2.320)
µt (ω) ¯ µt (ω) k3 + µi (ω) k30 ¯
and the fraction of the incident intensity which is transmitted is
¯ ¯2
k30 /µt (ω) + c.c. ¯¯ 2µt (ω) k3 ¯
¯ .
T⊥ = (2.321)
k3 /µi (ω) + c.c. µt (ω) k3 + µi (ω) k3 ¯
¯ 0
In the case that k3 and the permeabilities are real we have that
2µi (ω) µτ (ω) k3 (k30 + c.c.)
1 − R⊥ = = T⊥ (2.322)
|µt (ω) k3 + µi (ω) k30 |2
If k3 is complex (as is generally true) an interference between incident and reected waves occurs.
.................................................................................................................
.................................................................................................................
Assignment 11: The material of a thin film, in air, has a permittivity given by ε = 10 + 0.5i. (Assume the permeabilities
are one.) Light with a wave number k0 (real) is directed at the film at normal incidence. (a) By explicit calculation of each
part evaluate the fraction of the intensity which is (1) reected, (2) transmitted, (3) absorbed for film thickness d satisfying
0 < k0 d < 2π. (b) Repeat part (a) with ε = 10 − 0.5i. (−absorption is creation) (c) Compare and discuss the structures seen
in plots of your results.
.................................................................................................................
.................................................................................................................
2.7.2 Electric field parallel to the plane of incidence With the electric field parallel to the plane of
incidence the magnetic fields will be perpendicular to this plane. The magnetic field of the incident wave has an amplitude
Hi = Hi e2 (2.323)
and the magnetic fields of the reected and transmitted waves have amplitudes
Hr = Hr e2 (2.324)
Ht = Ht e2 . (2.325)
The fourier transform of the source free AmpereMaxwell law relates the electric and magnetic fields
ω
k × H = −ε (ω) E (2.326)
c
This gives the components of the electric fields which are parallel to e1 ,
ck3
(Ei )1 = Hi (2.327)
ωεi (ω)
ck3
(Er )1 = − Hr (2.328)
ωεi (ω)
ck30
(Et )1 = Ht (2.329)
ωεt (ω)
72
Section 2.7 Reection and refraction at a plane interface between dielectrics
2.7.3 Systems with multiple surfaces In the case that the electromagnetic wave passes through multiple
surfaces there will generally be Both incident and reected waves on each side of the interfaces. The most common example
of these systems involve the thin film coating of optical elements. The problem is to design and manufacture a coating which
will provide a given reectance or transmittance between elements. The importance of the problem is reected in the large
number of formulations of the problem and the multitude of computer codes designed to solve the problem.
73