Sie sind auf Seite 1von 18

2

1
Maxwells Equations

1. Maxwells Equations

the receiving antennas. Away from the sources, that is, in source-free regions of space,
Maxwells equations take the simpler form:

E=
H=

B
t

D
t

(source-free Maxwells equations)

(1.1.2)

D=0
B=0
The qualitative mechanism by which Maxwells equations give rise to propagating
electromagnetic elds is shown in the gure below.

1.1 Maxwells Equations


Maxwells equations describe all (classical) electromagnetic phenomena:

E=

B
t

H=J+

D
t

(Maxwells equations)

(1.1.1)

D=
B=0
The rst is Faradays law of induction, the second is Amp`
eres law as amended by
Maxwell to include the displacement current D/t, the third and fourth are Gauss laws
for the electric and magnetic elds.
The displacement current term D/t in Amp`
eres law is essential in predicting the
existence of propagating electromagnetic waves. Its role in establishing charge conservation is discussed in Sec. 1.7.
Eqs. (1.1.1) are in SI units. The quantities E and H are the electric and magnetic
eld intensities and are measured in units of [volt/m] and [ampere/m], respectively.
The quantities D and B are the electric and magnetic ux densities and are in units of
[coulomb/m2 ] and [weber/m2 ], or [tesla]. D is also called the electric displacement, and
B, the magnetic induction.
The quantities and J are the volume charge density and electric current density
(charge ux) of any external charges (that is, not including any induced polarization
charges and currents.) They are measured in units of [coulomb/m3 ] and [ampere/m2 ].
The right-hand side of the fourth equation is zero because there are no magnetic monopole charges. Eqs. (1.3.17)(1.3.19) display the induced polarization terms explicitly.
The charge and current densities , J may be thought of as the sources of the electromagnetic elds. For wave propagation problems, these densities are localized in space;
for example, they are restricted to ow on an antenna. The generated electric and magnetic elds are radiated away from these sources and can propagate to large distances to

For example, a time-varying current J on a linear antenna generates a circulating


and time-varying magnetic eld H, which through Faradays law generates a circulating
electric eld E, which through Amp`
eres law generates a magnetic eld, and so on. The
cross-linked electric and magnetic elds propagate away from the current source. A
more precise discussion of the elds radiated by a localized current distribution is given
in Chap. 14.

1.2 Lorentz Force


The force on a charge q moving with velocity v in the presence of an electric and magnetic eld E, B is called the Lorentz force and is given by:
F = q(E + v B)

(Lorentz force)

(1.2.1)

Newtons equation of motion is (for non-relativistic speeds):

dv
= F = q(E + v B)
dt

(1.2.2)

where m is the mass of the charge. The force F will increase the kinetic energy of the
charge at a rate that is equal to the rate of work done by the Lorentz force on the charge,
that is, v F. Indeed, the time-derivative of the kinetic energy is:

Wkin =

1
mv v
2

dWkin
dv
= mv
= v F = qv E
dt
dt

(1.2.3)

We note that only the electric force contributes to the increase of the kinetic energy
the magnetic force remains perpendicular to v, that is, v (v B)= 0.

1.3. Constitutive Relations

Volume charge and current distributions , J are also subjected to forces in the
presence of elds. The Lorentz force per unit volume acting on , J is given by:
f = E + J B

(Lorentz force per unit volume)

1. Maxwells Equations

The next simplest form of the constitutive relations is for simple homogeneous
isotropic dielectric and for magnetic materials:

(1.2.4)

D = E

where f is measured in units of [newton/m ]. If J arises from the motion of charges


within the distribution , then J = v (as explained in Sec. 1.6.) In this case,
f = (E + v B)

(1.2.5)

These are typically valid at low frequencies. The permittivity  and permeability
are related to the electric and magnetic susceptibilities of the material as follows:

By analogy with Eq. (1.2.3), the quantity v f = v E = J E represents the power


per unit volume of the forces acting on the moving charges, that is, the power expended
by (or lost from) the elds and converted into kinetic energy of the charges, or heat. It
has units of [watts/m3 ]. We will denote it by:

dPloss
=JE
dV

(ohmic power losses per unit volume)

(1.2.6)

(1.3.4)

B = H

 = 0 (1 + )

(1.3.5)

= 0 (1 + m )

The susceptibilities , m are measures of the electric and magnetic polarization


properties of the material. For example, we have for the electric ux density:
D = E = 0 (1 + )E = 0 E + 0 E = 0 E + P

In Sec. 1.8, we discuss its role in the conservation of energy. We will nd that electromagnetic energy owing into a region will partially increase the stored energy in that
region and partially dissipate into heat according to Eq. (1.2.6).

where the quantity P = 0 E represents the dielectric polarization of the material, that
is, the average electric dipole moment per unit volume. In a magnetic material, we have
B = 0 (H + M)= 0 (H + m H)= 0 (1 + m )H = H

1.3 Constitutive Relations


The electric and magnetic ux densities D, B are related to the eld intensities E, H via
the so-called constitutive relations, whose precise form depends on the material in which
the elds exist. In vacuum, they take their simplest form:
D = 0 E

(1.3.1)

B = 0 H

0 = 8.854 1012 farad/m


0 = 4 107 henry/m

(1.3.2)

The units for 0 and 0 are the units of the ratios D/E and B/H, that is,
coulomb
farad
coulomb/m2
=
=
,
volt/m
volt m
m

weber/m2
weber
henry
=
=
ampere/m
ampere m
m

From the two quantities 0 , 0 , we can dene two other physical constants, namely,
the speed of light and the characteristic impedance of vacuum:
1

c0 =
= 3 108 m/sec ,
0 0


0 =

0
= 377 ohm
0

(1.3.7)

where M = m H is the magnetization, that is, the average magnetic moment per unit
volume. The speed of light in the material and the characteristic impedance are:
1

c=


,

(1.3.8)

The relative permittivity, permeability and refractive index of a material are dened by:

rel =

where 0 , 0 are the permittivity and permeability of vacuum, with numerical values:

(1.3.6)


= 1 + ,
0

rel =

= 1 + m ,
0

n=

rel rel

(1.3.9)

so that n2 = rel rel . Using the denition of Eq. (1.3.8), we may relate the speed of light
and impedance of the material to the corresponding vacuum values:
1
c0
c0
=
=
=
0 0 rel rel
rel rel
n

 


0 rel
rel
rel
n
=
= 0
= 0
= 0
=

0
rel
rel
n
rel
1

c=

(1.3.10)

For a non-magnetic material, we have = 0 , or, rel = 1, and the impedance


becomes simply = 0 /n, a relationship that we will use extensively in this book.
More generally, constitutive relations may be inhomogeneous, anisotropic, nonlinear, frequency dependent (dispersive), or all of the above. In inhomogeneous materials,
the permittivity  depends on the location within the material:

(1.3.3)
D(r, t)= (r)E(r, t)

1.3. Constitutive Relations

In anisotropic materials,  depends on the x, y, z direction and the constitutive relations may be written component-wise in matrix (or tensor) form:


Dx
xx


Dy = yx
Dz
zx

xy
yy
zy

Ex
xz

yz Ey
zz
Ez

1. Maxwells Equations

In Sections 1.101.15, we discuss simple models of () for dielectrics, conductors,


and plasmas, and clarify the nature of Ohms law:
J = E

(1.3.11)

Anisotropy is an inherent property of the atomic/molecular structure of the dielectric. It may also be caused by the application of external elds. For example, conductors
and plasmas in the presence of a constant magnetic eldsuch as the ionosphere in the
presence of the Earths magnetic eldbecome anisotropic (see for example, Problem
1.10 on the Hall effect.)
In nonlinear materials,  may depend on the magnitude E of the applied electric eld
in the form:
D = (E)E ,
where
(E)=  + 2 E + 3 E2 +
(1.3.12)
Nonlinear effects are desirable in some applications, such as various types of electrooptic effects used in light phase modulators and phase retarders for altering polarization. In other applications, however, they are undesirable. For example, in optical bers
nonlinear effects become important if the transmitted power is increased beyond a few
milliwatts. A typical consequence of nonlinearity is to cause the generation of higher
harmonics, for example, if E = E0 ejt , then Eq. (1.3.12) gives:

D(r, t)=

(t t )E(r, t ) dt

D = 0 E + P ,

B = 0 (H + M)

(1.3.16)

Inserting these in Eq. (1.1.1), for example, by writing B = 0 (H + M)=


+ M), we may express Maxwells equations in
+ M)= 0 (0 E
+J+P
0 (J + D
terms of the elds E and B :

E=

B
t

B = 0 0
E=



E
P
+ 0 J +
+ M
t
t

0

(1.3.17)

P)

B=0
We identify the current and charge densities due to the polarization of the material as:

(1.3.13)
Jpol =

which becomes multiplicative in the frequency domain:


D(r, )= ()E(r, )

(1.3.15)

In Sec. 1.17, we discuss the Kramers-Kronig dispersion relations, which are a direct
consequence of the causality of the time-domain dielectric response function (t).
One major consequence of material dispersion is pulse spreading, that is, the progressive widening of a pulse as it propagates through such a material. This effect limits
the data rate at which pulses can be transmitted. There are other types of dispersion,
such as intermodal dispersion in which several modes may propagate simultaneously,
or waveguide dispersion introduced by the conning walls of a waveguide.
There exist materials that are both nonlinear and dispersive that support certain
types of non-linear waves called solitons, in which the spreading effect of dispersion is
exactly canceled by the nonlinearity. Therefore, soliton pulses maintain their shape as
they propagate in such media [1280,905,903].
More complicated forms of constitutive relationships arise in chiral and gyrotropic
media and are discussed in Chap. 4. The more general bi-isotropic and bi-anisotropic
media are discussed in [30,95]; see also [57].
In Eqs. (1.1.1), the densities , J represent the external or free charges and currents
in a material medium. The induced polarization P and magnetization M may be made
explicit in Maxwells equations by using the constitutive relations:

D = (E)E = E + 2 E2 + 3 E3 + = E0 ejt + 2 E02 e2jt + 3 E03 e3jt +


Thus the input frequency is replaced by , 2, 3, and so on. In a multiwavelength transmission system, such as a wavelength division multiplexed (WDM) optical ber system carrying signals at closely-spaced carrier frequencies, such nonlinearities will cause the appearance of new frequencies which may be viewed as crosstalk
among the original channels. For example, if the system carries frequencies i , i =
1, 2, . . . , then the presence of a cubic nonlinearity E3 will cause the appearance of the
frequencies i j k . In particular, the frequencies i + j k are most likely
to be confused as crosstalk because of the close spacing of the carrier frequencies.
Materials with a frequency-dependent dielectric constant () are referred to as
dispersive. The frequency dependence comes about because when a time-varying electric eld is applied, the polarization response of the material cannot be instantaneous.
Such dynamic response can be described by the convolutional (and causal) constitutive
relationship:

(Ohms law)

(1.3.14)

All materials are, in fact, dispersive. However, () typically exhibits strong dependence on only for certain frequencies. For example, water at optical frequencies has

refractive index n = rel = 1.33, but at RF down to dc, it has n = 9.

P
,
t

P
pol =

(polarization densities)

(1.3.18)

Similarly, the quantity Jmag = M may be identied as the magnetization current


density (note that mag = 0.) The total current and charge densities are:
Jtot = J + Jpol + Jmag = J +

tot = + pol = P

P
+ M
t

(1.3.19)

1.4. Negative Index Media

and may be thought of as the sources of the elds in Eq. (1.3.17). In Sec. 14.6, we examine
this interpretation further and show how it leads to the Ewald-Oseen extinction theorem
and to a microscopic explanation of the origin of the refractive index.

1.4 Negative Index Media

1. Maxwells Equations

current density; the difference of the normal components of the ux density D are equal
to the surface charge density; and the normal components of the magnetic ux density
B are continuous.
The Dn boundary condition may also be written a form that brings out the dependence on the polarization surface charges:

(0 E1n + P1n )(0 E2n + P2n )= s


Maxwells equations do not preclude the possibility that one or both of the quantities
, be negative. For example, plasmas below their plasma frequency, and metals up to
optical frequencies, have  < 0 and > 0, with interesting applications such as surface
plasmons (see Sec. 8.5).
Isotropic media with < 0 and  > 0 are more difcult to come by [155], although
examples of such media have been fabricated [383].
Negative-index media, also known as left-handed media, have , that are simultaneously negative,  < 0 and < 0. Veselago [378] was the rst to study their unusual
electromagnetic properties, such as having a negative index of refraction and the reversal of Snels law.
The novel properties of such media and their potential applications have generated
a lot of research interest [378460]. Examples of such media, termed metamaterials,
have been constructed using periodic arrays of wires and split-ring resonators, [384]
and by transmission line elements [417419,439,452], and have been shown to exhibit
the properties predicted by Veselago.
When rel < 0 and rel < 0, the refractive index, n2 = rel rel , must be dened by

the negative square root n = rel rel . Because then n < 0 and rel < 0 will imply
that the characteristic impedance of the medium = 0 rel /n will be positive, which
as we will see later implies that the energy ux of a wave is in the same direction as the
direction of propagation. We discuss such media in Sections 2.13, 7.16, and 8.6.

0 (E1n E2n )= s P1n + P2n = s,tot

The total surface charge density will be s,tot = s +1s,pol +2s,pol , where the surface
charge density of polarization charges accumulating at the surface of a dielectric is seen
to be (
n is the outward normal from the dielectric):
P
s,pol = Pn = n

(1.5.2)

The relative directions of the eld vectors are shown in Fig. 1.5.1. Each vector may
be decomposed as the sum of a part tangential to the surface and a part perpendicular
to it, that is, E = Et + En . Using the vector identity,
(E n
)+n
(n
E)= Et + En
E=n

(1.5.3)

we identify these two parts as:


(E n
) ,
Et = n

(n
E)= n
En
En = n

1.5 Boundary Conditions


The boundary conditions for the electromagnetic elds across material boundaries are
given below:
Fig. 1.5.1 Field directions at boundary.

E1t E2t = 0
H1t H2t

= Js n

D1n D2n = s

(1.5.1)

B1n B2n = 0

Using these results, we can write the rst two boundary conditions in the following
vectorial forms, where the second form is obtained by taking the cross product of the
and noting that Js is purely tangential:
rst with n
(E1 n
) n
(E2 n
) = 0
n
(H1 n
) n
(H2 n
) = Js n

is a unit vector normal to the boundary pointing from medium-2 into medium-1.
where n
The quantities s , Js are any external surface charge and surface current densities on
the boundary surface and are measured in units of [coulomb/m2 ] and [ampere/m].
In words, the tangential components of the E-eld are continuous across the interface; the difference of the tangential components of the H-eld are equal to the surface

or,

(E1 E2 ) = 0
n
(H1 H2 ) = Js
n

(1.5.4)

The boundary conditions (1.5.1) can be derived from the integrated form of Maxwells
equations if we make some additional regularity assumptions about the elds at the
interfaces.

1.6. Currents, Fluxes, and Conservation Laws

10

1. Maxwells Equations

In many interface problems, there are no externally applied surface charges or currents on the boundary. In such cases, the boundary conditions may be stated as:
E1t = E2t
H1t = H2t

D1n = D2n

(source-free boundary conditions)

(1.5.5)
Fig. 1.6.1 Flux of a quantity.

B1n = B2n

1.6 Currents, Fluxes, and Conservation Laws


The electric current density J is an example of a ux vector representing the ow of the
electric charge. The concept of ux is more general and applies to any quantity that
ows. It could, for example, apply to energy ux, momentum ux (which translates
into pressure force), mass ux, and so on.
In general, the ux of a quantity Q is dened as the amount of the quantity that
ows (perpendicularly) through a unit surface in unit time. Thus, if the amount Q
ows through the surface S in time t, then:

J=

Q
St

(denition of ux)

(1.6.1)

When the owing quantity Q is the electric charge, the amount of current through
the surface S will be I = Q/t, and therefore, we can write J = I/S, with units
of [ampere/m2 ].
The ux is a vectorial quantity whose direction points in the direction of ow. There
is a fundamental relationship that relates the ux vector J to the transport velocity v
and the volume density of the owing quantity:
J = v

(1.6.2)

This can be derived with the help of Fig. 1.6.1. Consider a surface S oriented perpendicularly to the ow velocity. In time t, the entire amount of the quantity contained
in the cylindrical volume of height vt will manage to ow through S. This amount is
equal to the density of the material times the cylindrical volume V = S(vt), that
is, Q = V = S vt. Thus, by denition:

S vt
Q
=
= v
J=
St
St

1.7 Charge Conservation


Maxwell added the displacement current term to Amp`
eres law in order to guarantee
charge conservation. Indeed, taking the divergence of both sides of Amp`
eres law and
using Gausss law D = , we get:

H = J+

(1.6.3)

In this sense, the terms electric and magnetic ux densities for the quantities D, B are somewhat of a
misnomer because they do not represent anything that ows.

=J+
D=J+
t
t
t

H = 0, we obtain the differential form of the charge


Using the vector identity
conservation law:

+ J = 0
t

(charge conservation)

(1.7.1)

Integrating both sides over a closed volume V surrounded by the surface S, as


shown in Fig. 1.7.1, and using the divergence theorem, we obtain the integrated form of
Eq. (1.7.1):


S

When J represents electric current density, we will see in Sec. 1.12 that Eq. (1.6.2)
implies Ohms law J = E. When the vector J represents the energy ux of a propagating
electromagnetic wave and the corresponding energy per unit volume, then because the
speed of propagation is the velocity of light, we expect that Eq. (1.6.2) will take the form:

Jen = cen

Similarly, when J represents momentum ux, we expect to have Jmom = cmom .


Momentum ux is dened as Jmom = p/(St)= F/S, where p denotes momentum and F = p/t is the rate of change of momentum, or the force, exerted on the
surface S. Thus, Jmom represents force per unit area, or pressure.
Electromagnetic waves incident on material surfaces exert pressure (known as radiation pressure), which can be calculated from the momentum ux vector. It can be
shown that the momentum ux is numerically equal to the energy density of a wave, that
is, Jmom = en , which implies that en = mom c. This is consistent with the theory of
relativity, which states that the energy-momentum relationship for a photon is E = pc.

J dS =

d
dt

dV

(1.7.2)

The left-hand side represents the total amount of charge owing outwards through
the surface S per unit time. The right-hand side represents the amount by which the
charge is decreasing inside the volume V per unit time. In other words, charge does not
disappear into (or created out of) nothingnessit decreases in a region of space only
because it ows into other regions.
Another consequence of Eq. (1.7.1) is that in good conductors, there cannot be any
accumulated volume charge. Any such charge will quickly move to the conductors
surface and distribute itself such that to make the surface into an equipotential surface.

1.8. Energy Flux and Energy Conservation

11

12

1. Maxwells Equations

where we introduce a change in notation:

en = w =

1
1
|E|2 + |H|2 = energy per unit volume
2
2

(1.8.2)

Jen = P = E H = energy ux or Poynting vector


where |E|2 = E E . The quantities w and P are measured in units of [joule/m3 ] and
[watt/m2 ]. Using the identity (E H)= H E E H, we nd:

Fig. 1.7.1 Flux outwards through surface.

w
H
E
+ P = 
E+
H + (E H)
t
t
t

Assuming that inside the conductor we have D = E and J = E, we obtain

B
D
E+
H+H EE H
t
t




D
B
=
H E+
+ E H
t
t
=

E= D=
J =



+ =0
t


(1.7.3)

with solution:

w
+ P = J E
t

(r, t)= 0 (r)et/


where 0 (r) is the initial volume charge distribution. The solution shows that the volume charge disappears from inside and therefore it must accumulate on the surface of
the conductor. The relaxation time constant rel = / is extremely short for good
conductors. For example, in copper,

rel

By contrast, rel is of the order of days in a good dielectric. For good conductors, the
above argument is not quite correct because it is based on the steady-state version of
Ohms law, J = E, which must be modied to take into account the transient dynamics
of the conduction charges.
It turns out that the relaxation time rel is of the order of the collision time, which
is typically 1014 sec. We discuss this further in Sec. 1.13. See also Refs. [138141].

(energy conservation)

(1.8.3)

As we discussed in Eq. (1.2.6), the quantity JE represents the ohmic losses, that
is, the power per unit volume lost into heat from the elds. The integrated form of
Eq. (1.8.3) is as follows, relative to the volume and surface of Fig. 1.7.1:

8.85 1012

=
=
= 1.6 1019 sec

5.7 107

P dS =

d
dt

w dV +

J E dV

(1.8.4)

It states that the total power entering a volume V through the surface S goes partially
into increasing the eld energy stored inside V and partially is lost into heat.
Example 1.8.1: Energy concepts can be used to derive the usual circuit formulas for capacitance, inductance, and resistance. Consider, for example, an ordinary plate capacitor with
plates of area A separated by a distance l, and lled with a dielectric . The voltage between
the plates is related to the electric eld between the plates via V = El.
The energy density of the electric eld between the plates is w = E2 /2. Multiplying this
by the volume between the plates, Al, will give the total energy stored in the capacitor.
Equating this to the circuit expression CV2 /2, will yield the capacitance C:

1.8 Energy Flux and Energy Conservation


Because energy can be converted into different forms, the corresponding conservation
equation (1.7.1) should have a non-zero term in the right-hand side corresponding to
the rate by which energy is being lost from the elds into other forms, such as heat.
Thus, we expect Eq. (1.7.1) to have the form:

en
+ Jen = rate of energy loss
t

Using Amp`
eres and Faradays laws, the right-hand side becomes:

(1.8.1)

Assuming the ordinary constitutive relations D = E and B = H, the quantities


en , Jen describing the energy density and energy ux of the elds are dened as follows,

W=

1 2
1
1
E Al = CV2 = CE2 l2
2
2
2

C=

A
l

Next, consider a solenoid with n turns wound around a cylindrical iron core of length
l, cross-sectional area A, and permeability . The current through the solenoid wire is
related to the magnetic eld in the core through Amp`
eres law Hl = nI. It follows that the
stored magnetic energy in the solenoid will be:

W=

1
1
1 H2 l2
H2 Al = LI2 = L 2
2
2
2
n

L = n2

A
l

Finally, consider a resistor of length l, cross-sectional area A, and conductivity . The


voltage drop across the resistor is related to the electric eld along it via V = El. The

1.9. Harmonic Time Dependence

13

current is assumed to be uniformly distributed over the cross-section A and will have
density J = E.
The power dissipated into heat per unit volume is JE = E2 . Multiplying this by the
resistor volume Al and equating it to the circuit expression V2 /R = RI2 will give:

(J E)(Al)= E2 (Al)=

E 2 l2
V2
=
R
R

R=

1 l
A

The same circuit expressions can, of course, be derived more directly using Q = CV, the


magnetic ux = LI, and V = RI.

Conservation laws may also be derived for the momentum carried by electromagnetic
elds [41,1243]. It can be shown (see Problem 1.6) that the momentum per unit volume
carried by the elds is given by:
G=DB=

c2

EH=
2

(momentum density)

14

1. Maxwells Equations

Next, we review some conventions regarding phasors and time averages. A realvalued sinusoid has the complex phasor representation:

A(t)= |A| cos(t + ) 




where A = |A|ej . Thus, we have A(t)= Re A(t) = Re Aejt . The time averages of
the quantities A(t) and A(t) over one period T = 2/ are zero.


The time average of the product of two harmonic quantities A(t)= Re Aejt and
jt
B(t)= Re Be
with phasors A, B is given by (see Problem 1.4):
A(t)B(t) =

A2 (t) =

E(r, )ejt

d
2

E(r, t)= E(r)ejt ,

H(r, t)= H(r)ejt

where the phasor amplitudes E(r), H(r) are complex-valued. Replacing time derivatives
by t j, we may rewrite Eq. (1.1.1) in the form:

E = jB
H = J + jD
D=


1
Re AB ]
2

(1.9.4)


1
1
Re AA ]= |A|2
2
2

(1.9.5)

A(t)B(t) dt =

T
0

A2 (t) dt =

1
1
1
E E + H H
Re
2
2
2


1

P = Re E H
2


dPloss
1
= Re Jtot E
dV
2

w=

(1.9.1)

Thus, we assume that all elds have a time dependence ejt :

Some interesting time averages in electromagnetic wave problems are the time averages of the energy density, the Poynting vector (energy ux), and the ohmic power
losses per unit volume. Using the denition (1.8.2) and the result (1.9.4), we have for
these time averages:

Maxwells equations simplify considerably in the case of harmonic time dependence.


Through the inverse Fourier transform, general solutions of Maxwells equation can be
built as linear combinations of single-frequency solutions:

In particular, the mean-square value is given by:

1.9 Harmonic Time Dependence

E(r, t)=

(1.9.2)

B=0
In this book, we will consider the solutions of Eqs. (1.9.2) in three different contexts:
(a) uniform plane waves propagating in dielectrics, conductors, and birefringent media, (b) guided waves propagating in hollow waveguides, transmission lines, and optical
bers, and (c) propagating waves generated by antennas and apertures.
The ejt convention is used in the engineering literature, and eit in the physics literature. One can
pass from one convention to the other by making the formal substitution j i in all the equations.

(energy density)
(Poynting vector)

(1.9.6)

(ohmic losses)

eres law and


where Jtot = J + jD is the total current in the right-hand side of Amp`
accounts for both conducting and dielectric losses. The time-averaged version of Poyntings theorem is discussed in Problem 1.5.
The expression (1.9.6) for the energy density w was derived under the assumption
that both  and were constants independent of frequency. In a dispersive medium, ,
become functions of frequency. In frequency bands where (), () are essentially
real-valued, that is, where the medium is lossless, it can be shown [155] that the timeaveraged energy density generalizes to:

(Maxwells equations)

(1.9.3)

(1.8.5)

where we set D = E, B = H, and c = 1/ . The quantity Jmom = cG = P /c will


represent momentum ux, or pressure, if the elds are incident on a surface.

A(t)= Aejt

w=

1
1 d()
1 d()
Re
E E +
H H
2
2 d
2 d


(lossless case)

(1.9.7)

The derivation of (1.9.7) is as follows. Starting with Maxwells equations (1.1.1) and
without assuming any particular constitutive relations, we obtain:
H B J E
E H = E D

(1.9.8)

As in Eq. (1.8.3), we would like to interpret the rst two terms in the right-hand side
as the time derivative of the energy density, that is,

dw
+ H B
=ED
dt

1.9. Harmonic Time Dependence

15

Anticipating a phasor-like representation, we may assume complex-valued elds and


derive also the following relationship from Maxwells equations:




1
1

1
1
Re H B Re J E
Re E H = Re E D
2
2
2
2

(1.9.9)

(1.9.10)

In a dispersive dielectric, the constitutive relation between D and E can be written


as follows in the time and frequency domains:

D(t)=

(t t )E(t )dt

D()= ()E()

where the Fourier transforms are dened by

(t)=

1
2

()ejt d

The time-derivative of D(t) is then


(t)=
D

where it follows from Eq. (1.9.12) that

(t)=

1
2

()=

(t)ejt dt

(t t )E(t )dt

j()ejt d

(1.9.11)

(1.9.12)

(1.9.13)

(1.9.14)

Following [155], we assume a quasi-harmonic representation for the electric eld,


E(t)= E 0 (t)ej0 t , where E 0 (t) is a slowly-varying function of time. Equivalently, in the
frequency domain we have E()= E 0 ( 0 ), assumed to be concentrated in a small
neighborhood of 0 , say, | 0 | . Because () multiplies the narrowband
function E(), we may expand () in a Taylor series around 0 and keep only the
linear terms, that is, inside the integral (1.9.14), we may replace:

()= a0 + b0 ( 0 ) ,

a0 = 0 (0 ) ,



d () 


b0 =

d
0

(1.9.15)

Inserting this into Eq. (1.9.14), we obtain the approximation

(t)

1
2


ja0 + b0 (j j0 ) ejt d = ja0 (t)+b0 (t j0 )(t) (1.9.16)

where (t) the Dirac delta function. This approximation is justied only insofar as it is
used inside Eq. (1.9.13). Inserting (1.9.16) into Eq. (1.9.13), we nd
(t) =
D


ja0 (t t )+b0 (t j0 )(t t ) E(t )dt =

= ja0 E(t)+b0 (t j0 )E(t)




= ja0 E 0 (t)ej0 t + b0 (t j0 ) E 0 (t)ej0 t


= ja0 E 0 (t)+b0 E0 (t) ej0 t
To

unclutter the notation, we are suppressing the dependence on the space coordinates r.

1. Maxwells Equations

Because we assume that () is real (i.e., lossless) in the vicinity of 0 , it follows that:


 1


1

1
= Re E 0 (t) ja0 E 0 (t)+b0 E0 (t) = b0 Re E 0 (t) E0 (t) ,
Re E D
2
2
2


d
1
=
Re E D
2
dt

from which we may identify a time-averaged version of dw/dt,

dw
1
+ Re H B
= Re E D
dt
2
2

16

(1.9.17)

1
b0 |E 0 (t)|2
4


=

d
dt



1 d () 0
|E 0 (t)|2
4
d

or,

(1.9.18)

Dropping the subscript 0, we see that the quantity under the time derivative in the
right-hand side may be interpreted as a time-averaged energy density for the electric
eld. A similar argument can be given for the magnetic energy term of Eq. (1.9.7).
We will see in the next section that the energy density (1.9.7) consists of two parts:
one part is the same as that in the vacuum case; the other part arises from the kinetic
and potential energy stored in the polarizable molecules of the dielectric medium.
When Eq. (1.9.7) is applied to a plane wave propagating in a dielectric medium, one
can show that (in the lossless case) the energy velocity coincides with the group velocity.
The generalization of these results to the case of a lossy medium has been studied
extensively [155169]. Eq. (1.9.7) has also been applied to the case of a left-handed
medium in which both () and () are negative over certain frequency ranges. As
argued by Veselago [378], such media must necessarily be dispersive in order to make
Eq. (1.9.7) a positive quantity even though individually  and are negative.
Analogous expressions to (1.9.7) may also be derived for the momentum density of
a wave in a dispersive medium. In vacuum, the time-averaged momentum density is
given by Eq. (1.8.5), that is,


1
= Re 0 0 E H
G
2
For the dispersive (and lossless) case this generalizes to [378,454]

=
G

1
k
Re E H +
2
2

d
d
|E|2 +
|H|2
d
d


(1.9.19)

1.10 Simple Models of Dielectrics, Conductors, and Plasmas


A simple model for the dielectric properties of a material is obtained by considering the
motion of a bound electron in the presence of an applied electric eld. As the electric
eld tries to separate the electron from the positively charged nucleus, it creates an
electric dipole moment. Averaging this dipole moment over the volume of the material
gives rise to a macroscopic dipole moment per unit volume.
A simple model for the dynamics of the displacement x of the bound electron is as
= dx/dt):
follows (with x
= eE kx mx

mx
(1.10.1)
where we assumed that the electric eld is acting in the x-direction and that there is
a spring-like restoring force due to the binding of the electron to the nucleus, and a
friction-type force proportional to the velocity of the electron.
The spring constant k is related to the resonance frequency of the spring via the

relationship 0 = k/m, or, k = m20 . Therefore, we may rewrite Eq. (1.10.1) as


+ x
+ 20 x =
x

e
E
m

(1.10.2)

1.11. Dielectrics

17

The limit 0 = 0 corresponds to unbound electrons and describes the case of good
arises from collisions that tend to slow down the
conductors. The frictional term x
electron. The parameter is a measure of the rate of collisions per unit time, and
therefore, = 1/ will represent the mean-time between collisions.
In a typical conductor, is of the order of 1014 seconds, for example, for copper,
= 2.4 1014 sec and = 4.1 1013 sec1 . The case of a tenuous, collisionless,
plasma can be obtained in the limit = 0. Thus, the above simple model can describe
the following cases:
a. Dielectrics, 0 = 0, = 0.
b. Conductors, 0 = 0, = 0.
c. Collisionless Plasmas, 0 = 0, = 0.

18

1. Maxwells Equations

The electric ux density will be then:



D = 0 E + P = 0 1 + () E ()E
where the effective permittivity () is:

Ne2
m
()= 0 + 2
0 2 + j

(1.11.4)

This can be written in a more convenient form, as follows:

()= 0 +

The basic idea of this model is that the applied electric eld tends to separate positive
from negative charges, thus, creating an electric dipole moment. In this sense, the
model contains the basic features of other types of polarization in materials, such as
ionic/molecular polarization arising from the separation of positive and negative ions
by the applied eld, or polar materials that have a permanent dipole moment.

1.11 Dielectrics

0 2p
20

(1.11.5)

2 + j

where 2p is the so-called plasma frequency of the material dened by:

2p =

Ne2
0 m

(plasma frequency)

(1.11.6)

The model dened by (1.11.5) is known as a Lorentz dielectric. The corresponding




susceptibility, dened through ()= 0 1 + () , is:

The applied electric eld E(t) in Eq. (1.10.2) can have any time dependence. In particular,
if we assume it is sinusoidal with frequency , E(t)= Eejt , then, Eq. (1.10.2) will have
the solution x(t)= xejt , where the phasor x must satisfy:

2 x + jx + 20 x =

e
E
m

()=

2p
20

For a dielectric, we may assume 0 = 0. Then, the low-frequency limit ( = 0) of


Eq. (1.11.5), gives the nominal dielectric constant:

which is obtained by replacing time derivatives by t j. Its solution is:

e
E
m
x= 2
2
0 + j

(0)= 0 + 0
(1.11.1)

The corresponding velocity of the electron will also be sinusoidal v(t)= vejt , where
= jx. Thus, we have:
v=x

e
j E
m
v = jx = 2
0 2 + j

(1.11.2)

From Eqs. (1.11.1) and (1.11.2), we can nd the polarization per unit volume P.
We assume that there are N such elementary dipoles per unit volume. The individual
electric dipole moment is p = ex. Therefore, the polarization per unit volume will be:

Ne2
E
m
0 ()E
P = Np = Nex = 2
2
0 + j

(1.11.7)

2 + j

(1.11.3)

2p
20

= 0 +

Ne2
m20

(1.11.8)

The real and imaginary parts of () characterize the refractive and absorptive
properties of the material. By convention, we dene the imaginary part with the negative
sign (because we use ejt time dependence):

()=  ()j ()

(1.11.9)

It follows from Eq. (1.11.5) that:

 ()= 0 +

0 2p (20 2 )
(2

20 )2 +2 2

 ()=

0 2p
(2

20 )2 +2 2

(1.11.10)

Fig. 1.11.1 shows a plot of  () and  (). Around the resonant frequency 0 ,
the real part  () behaves in an anomalous manner, that is, it drops rapidly with
frequency to values less than 0 and the material exhibits strong absorption. The term
normal dispersion refers to an  () that is an increasing function of , as is the
case to the far left and right of the resonant frequency.

1.11. Dielectrics

19

20

1. Maxwells Equations

Fig. 1.11.1 Real and imaginary parts of the effective permittivity ().

Fig. 1.11.2 Effective permittivity in a two-level gain medium with f = 1.

Real dielectric materials exhibit, of course, several such resonant frequencies corresponding to various vibrational modes and polarization mechanisms (e.g., electronic,
ionic, etc.) The permittivity becomes the sum of such terms:

In practice, Eq. (1.11.14) is applied in frequency ranges that are far from any resonance so that one can effectively set i = 0:

()= 0 + 0

Ni ei /mi 0
i 2 + ji
2

n2 ()= 1 +

(1.11.11)

A more correct quantum-mechanical treatment leads essentially to the same formula:

()= 0 + 0

 fji (Ni Nj )e2 /m0


j>i

2ji 2 + jji

(1.11.12)

where ji are transition frequencies between energy levels, that is, ji = (Ej Ei )/,
and Ni , Nj are the populations of the lower, Ei , and upper, Ej , energy levels. The quantities fji are called oscillator strengths. For example, for a two-level atom we have:

()= 0 + 0

f = f21

20 2 + j

N1 N2
,
N1 + N2

2p =


i

(Sellmeier equation)

(1.11.15)

where , i denote the corresponding free-space wavelengths (e.g., = 2c/). In


practice, refractive index data are tted to Eq. (1.11.15) using 24 terms over a desired
frequency range. For example, fused silica (SiO2 ) is very accurately represented over the
range 0.2 3.7 m by the following formula [147], where and i are in units of
m:

n2 = 1 +

0.6961663 2
0.4079426 2
0.8974794 2
+ 2
+ 2
2
2
(0.0684043)
(0.1162414)
(9.896161)2

(1.11.16)

(1.11.13)
The conductivity properties of a material are described by Ohms law, Eq. (1.3.15). To
derive this law from our simple model, we use the relationship J = v, where the volume
density of the conduction charges is = Ne. It follows from Eq. (1.11.2) that

(N1 + N2 )e2
m0

Bi 2i
2i 2 + ji

Ne2
E
m
()E
J = v = Nev = 2
0 2 + j
j

Normally, lower energy states are more populated, Ni > Nj , and the material behaves
as a classical absorbing dielectric. However, if there is population inversion, Ni < Nj ,
then the corresponding permittivity term changes sign. This leads to a negative imaginary part,  (), representing a gain. Fig. 1.11.2 shows the real and imaginary parts
of Eq. (1.11.13) for the case of a negative effective oscillator strength f = 1.
The normal and anomalous dispersion bands still correspond to the bands where
the real part  () is an increasing or decreasing, respectively, function of frequency.
But now the normal behavior is only in the neighborhood of the resonant frequency,
whereas far from it,the behavior is anomalous.
Setting n()= ()/0 for the refractive index, Eq. (1.11.11) can be written in the
following form, known as the Sellmeier equation (where the Bi are constants):

n2 ()= 1 +

 Bi 2
Bi 2i
=1+
2
2
i
2 2i
i

1.12 Conductors

f 2p

where we dened:

0 = 21 ,

(1.11.14)

and therefore, we identify the conductivity ():

Ne2
j0 2p
m
= 2
()= 2
2
0 + j
0 2 + j
j

(1.12.1)

We note that ()/j is essentially the electric susceptibility considered above.


Indeed, we have J = Nev = Nejx = jP, and thus, P = J/j = (()/j)E. It
follows that ()0 = ()/j, and

()= 0 +

0 2p
20

2 + j

= 0 +

()
j

(1.12.2)

1.12. Conductors

21

Since in a metal the conduction charges are unbound, we may take 0 = 0 in


Eq. (1.12.1). After canceling a common factor of j , we obtain:

()=

0 2p

(1.12.3)

+ j

22

1. Maxwells Equations

where u(t) is the unit-step function. As an example, suppose the electric eld E(t) is a
constant electric eld that is suddenly turned on at t = 0, that is, E(t)= Eu(t). Then,
the time response of the current will be:

t
J(t)=
0

The model dened by (1.12.3) is know as the Drude model. The nominal conductivity is obtained at the low-frequency limit, = 0:

0 2p

Ne2
m

(nominal conductivity)

(1.12.4)

Example 1.12.1: Copper has a mass density of 8.9 106 gr/m3 and atomic weight of 63.54
(grams per mole.) Using Avogadros number of 6 1023 atoms per mole, and assuming
one conduction electron per atom, we nd for the volume density N:

where = 0 p / is the nominal conductivity of the material.


Thus, the current starts out at zero and builds up to the steady-state value of J = E,
which is the conventional form of Ohms law. The rise time constant is = 1/. We
saw above that is extremely smallof the order of 1014 secfor good conductors.
The building up of the current can also be understood in terms of the equation of
motion of the conducting charges. Writing Eq. (1.10.2) in terms of the velocity of the
charge, we have:
(t)+v(t)=
v

t
v(t)=

(8.4 1028 )(1.6 1019 )2


Ne2
=
= 5.8 107 Siemens/m
m
(9.1 1031 )(4.1 1013 )

where we used e = 1.6 1019 , m = 9.1 1031 , = 4.1 1013 . The plasma frequency
of copper can be calculated by

Ne2
= 2.6 1015 Hz
m0

which lies in the ultraviolet range. For frequencies such that  , the conductivity
(1.12.3) may be considered to be independent of frequency and equal to the dc value of
Eq. (1.12.4). This frequency range covers most present-day RF applications. For example,


assuming 0.1, we nd f 0.1/2 = 653 GHz.

So far, we assumed sinusoidal time dependence and worked with the steady-state
responses. Next, we discuss the transient dynamical response of a conductor subject to
an arbitrary time-varying electric eld E(t).
Ohms law can be expressed either in the frequency-domain or in the time-domain
with the help of the Fourier transform pair of equations:

t
J()= ()E()

J(t)=

(t t )E(t )dt

e(tt )


e
e
E(t )dt =
E 1 et
m
m

For large t, the velocity reaches the steady-state value v = (e/m)E, which reects
the balance between the accelerating electric eld force and the retarding frictional force,
that is, mv = eE. The quantity e/m is called the mobility of the conduction charges.
The steady-state current density results in the conventional Ohms law:

It follows that:

p
1
=
2
2

e
E(t)
m

Assuming E(t)= Eu(t), we obtain the convolutional solution:

atoms
6 10
mole 8.9 106 gr  1 electron  = 8.4 1028 electrons/m3
N=
gr
m3
atom
63.54
mole

fp =

0 2p



E 1 et = E 1 et

23

0 2p e(tt ) Edt =

J = Nev =

Ne2
E = E
m

A more accurate description of the permittivity properties of metals, especially at


optical and infrared frequencies which are relevant in plasmonic waveguides, requires
the addition of interband terms, generalizing the Drude model to the so-called DrudeLorentz model of the form,
k

2p
fi 2p
()
+
=1+
2
0
j( + j) i=1 i 2 + ji

(Drude-Lorentz model) (1.12.7)

Rakic, et al. [154] have tted 11 metals, such as silver, gold, aluminum, copper,
to such an expression with 56 terms, covering a wide range of frequencies and wavelengths, 25 THz < f < 1500 THz, or, equivalently, 200 nm < < 12 m. The MATLAB
function, drude, implements the results of [154],
ep = drude(lambda,metal)

% Drude-Lorentz model for Silver, Gold, Copper, Aluminum

(1.12.5)

where (t) is the causal inverse Fourier transform of (). For the simple model of
Eq. (1.12.3), we have:
(t)= 0 2p et u(t)
(1.12.6)

lambda
metal

= vector of wavelengths in nanometers


= s, g, c, a, for silver, gold, copper, aluminum

ep = complex relative permittivity (same size as lambda)

1.13. Charge Relaxation in Conductors

23

1.13 Charge Relaxation in Conductors

24

1. Maxwells Equations

Denoting the rst two terms by d () and the third by c ()/j, we obtain the
total effective permittivity of such a material:

Next, we discuss the issue of charge relaxation in good conductors [138141]. Writing
(1.12.5) three-dimensionally and using (1.12.6), Ohms law reads in the time domain:

(tt )

J(r, t)= p

0 E(r, t ) dt

(1.13.1)

= 0,
Taking the divergence of both sides and using charge conservation, J +
and Gausss law, 0 E = , we obtain the following integro-differential equation for
the charge density (r, t):

t


r, t)= J(r, t)= 2p
(
e(tt ) 0 E(r, t )dt = 2p
e(tt ) (r, t )dt

Differentiating both sides with respect to t, we nd that satises the second-order


differential equation:
r, t)+2p (r, t)= 0
r, t)+(
(
(1.13.2)
whose solution is easily veried to be a linear combination of:


et/2 cos(relax t) ,

et/2 sin(relax t) ,

where

relax =

2p

relax =

= 2

(relaxation time constant)

(1.13.3)

1.14 Power Losses


To describe a material with both dielectric and conductivity properties, we may take the
susceptibility to be the sum of two terms, one describing bound polarized charges and
the other unbound conduction charges. Assuming different parameters {0 , p , } for
each term, we obtain the total permittivity:

0 2dp
2

d0

+ jd

0 2cp
j(c + j)

(effective permittivity)

(1.14.2)

In the low-frequency limit, = 0, the quantities d (0) and c (0) represent the
nominal dielectric constant and conductivity of the material. We note also that we can
write Eq. (1.14.2) in the form:

j()= c ()+jd ()

(1.14.3)

These two terms characterize the relative importance of the conduction current and
the displacement (polarization) current. The right-hand side in Amp`
eres law gives the
total effective current:

Jtot = J +

D
= J + jD = c ()E + jd ()E = j()E
t



J

|c ()|
|c ()E|
 cond 

=
=
 Jdisp  |jd ()E|
|d ()|

Typically, p  , so that relax is practically equal to p . For example, using


the numerical data of Example 1.12.1, we nd for copper relax = 2 = 51014 sec.
We calculate also: frelax = relax /2 = 2.61015 Hz. In the limit , or 0,
Eq. (1.13.2) reduces to the naive relaxation equation (1.7.3) (see Problem 1.9).
In addition to charge relaxation, the total relaxation time depends on the time it takes
for the electric and magnetic elds to be extinguished from the inside of the conductor,
as well as the time it takes for the accumulated surface charge densities to settle, the
motion of the surface charges being damped because of ohmic losses. Both of these
times depend on the geometry and size of the conductor [140].

()= 0 +

c ()
j

where the term Jdisp = D/t = jd ()E represents the displacement current. The
relative strength between conduction and displacement currents is the ratio:

Thus, the charge density is an exponentially decaying sinusoid with a relaxation time
constant that is twice the collision time = 1/:
2

()= d ()+

(1.14.1)

(1.14.4)

This ratio is frequency-dependent and establishes a dividing line between a good


conductor and a good dielectric. If the ratio is much larger than unity (typically, greater
than 10), the material behaves as a good conductor at that frequency; if the ratio is much
smaller than one (typically, less than 0.1), then the material behaves as a good dielectric.
Example 1.14.1: This ratio can take a very wide range of values. For example, assuming a
frequency of 1 GHz and using (for illustration purposes) the dc-values of the dielectric
constants and conductivities, we nd:

9



10
J


 cond 
1

=
=

 Jdisp  
9
10

for copper with = 5.8107 S/m and  = 0


for seawater with = 4 S/m and  = 720
for a glass with = 1010 S/m and  = 20

Thus, the ratio varies over 18 orders of magnitude! If the frequency is reduced by a factor
of ten to 100 MHz, then all the ratios get multiplied by 10. In this case, seawater acts like


a good conductor.

The time-averaged ohmic power losses per unit volume within a lossy material are
given by Eq. (1.9.6). Writing ()=  ()j (), we have:
Jtot = j()E = j ()E +  ()E

 

2
Denoting E  = E E , it follows that:

 2

1
dPloss
1
= Re Jtot E =  ()E 
2
2
dV

(ohmic losses)

(1.14.5)

1.15. Plasmas

25

26

1. Maxwells Equations

Writing d ()= d ()j


d () and assuming that the conductivity c () is realvalued for the frequency range of interest (as was discussed in Example 1.12.1), we nd
by equating real and imaginary parts of Eq. (1.14.2):

 ()= d () ,

 ()= 


d ()+

c ()

It follows that for a plasma:


(1.14.6)

Then, the power losses can be written in a form that separates the losses due to
conduction and those due to the polarization properties of the dielectric:

 2
dPloss
1
E 
c ()+
=
d ()
dV
2

(ohmic losses)

(1.14.7)

A convenient way to quantify the losses is by means of the loss tangent dened in
terms of the real and imaginary parts of the effective permittivity:
tan =

 ()
 ()



 1
k = 0 0 1 2p /2 =
2 2p
(1.15.2)
c

where we used c = 1/ 0 0 .
If > p , the electromagnetic wave propagates without attenuation within the
plasma. But if < p , the wavenumber k becomes imaginary and the wave gets
attenuated. At such frequencies, a wave incident (normally) on the ionosphere from the
ground cannot penetrate and gets reected back.

1.16 Energy Density in Lossless Dispersive Dielectrics


(loss tangent)

(1.14.8)

where is the loss angle. Eq. (1.14.8) may be written as the sum of two loss tangents,
one due to conduction and one due to polarization. Using Eq. (1.14.6), we have:

tan =


k = ()

c ()+
 ()
c ()
d ()
=
+ d
= tan c + tan d


d ()
d ()
d ()

(1.14.9)

The lossless case is obtained from Eq. (1.11.5) by setting = 0, which is equivalent to
assuming that is far from the resonance 0 . In this case the permittivity is:

()= 0 1 +



2p (2 + 20 )
d()
= 0 1 +
d
(20 2 )2

(1.16.1)

Thus, the electric part of the energy density (1.9.7) will be:

 2
dPloss
1
= d ()tan E 
dV
2

(ohmic losses)

(1.14.10)
e =
w

1.15 Plasmas
To describe a collisionless plasma, such as the ionosphere, the simple model considered in the previous sections can be specialized by choosing 0 = = 0. Thus, the
conductivity given by Eq. (1.12.3) becomes pure imaginary:

2p (2 + 20 )
1 d()
1
|E|2 = 0 |E|2 1 +
4 d
4
(20 2 )2

x=

()
= 0 1 2
j

(1.16.2)

eE/m
,
20 2

v = jx

The time-averaged mechanical energy (per unit volume) is obtained by adding the
kinetic and potential energies:

mech =
w


2

This expression can be given a nice interpretation: The rst term on the right is the
energy density in vacuum and the second corresponds to the mechanical (kinetic and
potential) energy of the polarization charges [156,179]. Indeed, the displacement x and
of the polarization charges are in this case:
velocity v = x

0 2p

The corresponding effective permittivity of Eq. (1.12.2) becomes purely real:

()= 0 +

20 2

from which it follows that:

The ohmic loss per unit volume can be expressed in terms of the loss tangent as:

()=

2p

(1.15.1)

2
= Ne2 /m0 . In the ionosphere
The plasma frequency can be calculated from p
the electron density is typically N = 1012 , which gives fp = 9 MHz.
We will see in Sec. 2.6 that the propagation wavenumber of an electromagnetic wave
propagating in a dielectric/conducting medium is given in terms of the effective permittivity by:

1
Re N
2

1
1
m|v|2 + m20 |x|2
2
2



1
Nm(2 + 20 )|x|2
4

1 Nm(2 + 20 )e2 |E|2 /m2


1
= 0 |E|2
4
4
(20 2 )2

2p (2 + 20 )

(20 2 )2

where we used the denition (1.11.6) of the plasma frequency. It follows that Eq. (1.16.2)
can be written as the sum:
e =
w

1 d()
1
mech = w
vac + w
mech
|E|2 = 0 |E|2 + w
4 d
4

(1.16.3)

1.17. Kramers-Kronig Dispersion Relations

27

1.17 Kramers-Kronig Dispersion Relations

(t t )E(r, t )dt =

(t t )E(r, t )dt

Because D(r, t)= 0 E(r, t)+P(r, t), we may dene the time-domain susceptibility
function (t) through:
(t)= 0 (t)+0 (t)
(1.17.1)
where (t) is the Dirac delta function. Therefore, if (t) is causal, so is (t). The
polarization is then given by:

t
P(r, t)= 0

(t t )E(r, t )dt = 0

(t t )E(r, t )dt

(1.17.2)

In the frequency domain, this becomes multiplicative: P(r, )= 0 ()E(r, ).


The Kramers-Kronig relations are the frequency-domain expression of causality and relate the real and imaginary parts of the susceptibility function (). Here, the functions
(t) and () are Fourier transform pairs:

()=

(t)ejt dt

(t)=

()ejt d

for all t

(1.17.4)

Using the property that the Fourier transform of a product of two time functions is
the convolution of their Fourier transforms, it follows that Eq. (1.17.4) can be written in
the equivalent frequency-domain form:
1
()=
2

0+

( )U( )d

j + 

=P

1
2j

+ ()

( ) P

j(

 )

+ ( ) d

( )
1
d + ()

2

(Kramers-Kronig)

(1.17.7)

The reason for applying this relation to () instead of () is that () falls off
sufciently fast for large to make the integral in (1.17.5) convergent, whereas ()
tends to the constant 0 .
Setting ()= r ()ji () and separating (1.17.7) into its real and imaginary
parts, we obtain the conventional form of the Kramers-Kronig dispersion relations:

r () =

i () =

P
1

i ( )
d


(Kramers-Kronig relations)

r ( )
d


(1.17.8)

Because the time-response (t) is real-valued, its Fourier transform () will satisfy the Hermitian symmetry property ()= (), which is equivalent to the even
symmetry of its real part, r ()= r (), and the odd symmetry of its imaginary part,
i ()= i (). Taking advantage of these symmetries, the range of integration in
(1.17.8) can be folded in half resulting in:

i () =

P
2

 i ( )
d
2 2

r ( )
d
2 2

(1.17.9)

There are several other ways to prove the Kramers-Kronig relations. For example,
a more direct way is to state the causality condition in terms of the signum function


sign(t). Indeed, because u(t)= 1 + sign(t) /2, Eq. (1.17.4) may be written in the
equivalent form (t)= (t)sign(t). Then, Eq. (1.17.7) follows by applying the same
frequency-domain convolution argument using the Fourier transform pair:

(1.17.10)

Alternatively, the causality condition can be expressed as u(t)(t)= 0. This approach is explored in Problem 1.12. Another proof is based on the analyticity properties
of (). Because of the causality condition, the Fourier integral in (1.17.3) can be restricted to the time range 0 < t < :

(1.17.6)

( )
d


(1.17.5)

where P denotes the principal value. Inserting (1.17.6) into (1.17.5), we have:
1
() =
2

sign(t) 

where U() is the Fourier transform of the unit-step. Eq. (1.17.5) is essentially the
Kramers-Kronig relation. The function U() is given by the well-known expression:

U()= lim

(1.17.3)

The causality condition, (t)= 0 for t < 0, can be expressed in terms of the unit-step
function u(t) in the equivalent manner:

(t)= (t)u(t) ,

()=

r () =

1
2

1. Maxwells Equations

Rearranging terms and canceling a factor of 1/2, we obtain the Kramers-Kronig relation in its complex-valued form:

The convolutional form of Eq. (1.3.13) implies causality, that is, the value of D(r, t) at
the present time t depends only on the past values of E(r, t ), t t.
This condition is equivalent to requiring that the dielectric response (t) be a rightsided (causal) function of time, that is, (t)= 0 for t < 0. Then, Eq. (1.3.13) may be
written as ordinary convolution by extending the integration range over all times:
D(r, t)=

28

()=

ejt (t)dt =

ejt (t)dt

(1.17.11)

This implies that () can be analytically continued into the lower half -plane,
so that replacing by w = j with 0 still gives a convergent Fourier integral
The right-hand side (without the j) in (1.17.7) is known as a Hilbert transform. Exchanging the roles
of t and , such transforms, known also as 90o phase shifters, are used widely in signal processing for
generating single-sideband communications signals.

1.18. Group Velocity, Energy Velocity

29

in Eq. (1.17.11). Any singularities in () lie in the upper-half plane. For


 example, the

20

0 =
0 + j/2, where
simple model of Eq. (1.11.7) has poles at =

Next, we consider a clockwise closed contour C = C + C consisting of the real axis


C and an innite semicircle C in the lower half-plane. Because () is analytic in the
region enclosed by C, Cauchys integral theorem implies that for any point w enclosed
by C, that is, lying in the lower half-plane, we must have:

(w)=

1
2j

(w )
dw
w w

2 /4.

(1.17.12)

30

1. Maxwells Equations

This form preserves the sign of i , that is, ni and i are both positive for absorbing
media, or both negative for gain media. The following approximate solution is often
used, which can be justied whenever ||  1 (for example, in gases):

nr ()jnr ()= 1 + () 1 +
2

nr = 1 +

1
r ,
2

ni =

1
i
2

We will see in Chap. 2 that a single-frequency uniform plane wave propagating, say,
in the positive z-direction, has a wavenumber k = n/c = (nr jni )/c j,
where c is the speed of light in vacuum. Therefore, the wave will have a space-time
dependence:

ej(tkz) = ej(t(j)z) = ez ej(tz) = eni z/c ej(tnr z/c)


where the overall minus sign arises because C was taken to be clockwise. Assuming that
() falls off sufciently fast for large , the contribution of the innite semicircle
can be ignored, thus leaving only the integral over the real axis. Setting w = j and
taking the limit  0+, we obtain the identical relationship to Eq. (1.17.5):

()= lim

0+

1
2j

( )
1
d =

+ j
2

( ) lim

0+

j(  )+

d

An interesting consequence of the Kramers-Kronig relations is that there cannot


exist a dielectric medium that is purely lossless, that is, such that i ()= 0 for all ,
because this would also require that r ()= 0 for all .
However, in all materials, i () is signicantly non-zero only in the neighborhoods
of the mediums resonant frequencies, as for example in Fig. 1.11.1. In the frequency
bands that are sufciently far from the resonant bands, i () may be assumed to be
essentially zero. Such frequency bands are called transparency bands [155].

1.18 Group Velocity, Energy Velocity


Assuming a nonmagnetic material ( = 0 ), a complex-valued refractive index may be
dened by:



n()= nr ()jni ()= 1 + () =

()
0

(1.18.1)

where nr , ni are its real and imaginary parts. Setting = r ji we have the condition

nr jni = 1 + r ji . Upon squaring, this splits into the two real-valued equations
n2r n2i = 1 + r and 2nr ni = i , with solutions:

1/2

(1 + r )2 +2i + (1 + r )

nr =
2


1/2
(1 + r )2 +2i (1 + r )
i

ni = sign(i )
=
2
2nr

(1.18.2)

(1.18.3)

(1.18.4)

The real part nr denes the phase velocity of the wave, vp = / = c/nr , whereas
the imaginary part ni , or = ni /c, corresponds to attenuation or gain depending on
the sign of ni or i .
When several such plane waves are superimposed to form a propagating pulse, we
will see in Sec. 3.5 that the peak of the pulse (i.e., the point on the pulse where all the
individual frequency components add up in phase), propagates with the so-called group
velocity dened by:

vg =

1
d
c
c
=
= group velocity
=
=
d(nr )
dnr
d
d
nr +
d
d
d

(1.18.5)

A group refractive index may be dened through vg = c/ng , or, ng = c/vg :

ng =

d(nr )
dnr
dnr
= nr +
= nr
= group refractive index
d
d
d

(1.18.6)

where is the free-space wavelength related to by = 2c/, and we used the


differentiation property that d/d = d/d.
Within an anomalous dispersion region, nr is decreasing rapidly with , that is,
dnr /d < 0, as in Fig. 1.11.1. This results in a group velocity vg , given by Eq. (1.18.5),
that may be larger than c or even negative. Such velocities are called superluminal.
Light pulses propagating at superluminal group velocities are referred to as fast light
and we discuss them further in Sec. 3.9.
Within a normal dispersion region (e.g., to the far left and far right of the resonant
frequency 0 in Fig. 1.11.1), nr is an increasing function of , dnr /d > 0, which
results in vg < c. In specially engineered materials such as those exhibiting electromagnetically induced transparency, the slope dnr /d may be made so steep that the
resulting group velocity vg becomes extremely small, vg  c. This is referred to as
slow light.
We close this section by showing that for lossless dispersive media, the energy velocity of a plane wave is equal to the group velocity dened by (1.18.5). This result is
quite general, regardless of the frequency dependence of () and () (as long as
these quantities are real.)

1.19. Problems

31

We will see in the next chapter that a plane wave propagating along the z-direction
has electric and magnetic elds that are transverse to the z-direction and are related by:

|H| =


|E| ,

Moreover the time-averaged energy ux (in the z-direction) and energy density are:
z =
P

|E|2
,
2

=
w

1 d()
1 d()
1
|E|2 +
|H|2 =
4 d
4 d
4

1 d()
d()
+ 2
d


|E|2

z /w
. Thus, we have:
The energy velocity is dened by ven = P
1
ven

w
1
=
=
z
2
P

1 d()
d()

+
d
d

1
=
2



d()
+
 d

 d()
d

It is easily veried that the right-hand side can be expressed in terms of the wave
number k =  in the form:
1

ven

1
=
2



d()
+
 d

 d()
d


d 
dk
=
=
= vg1
d
d

(1.18.7)

which shows the equality of the energy and group velocities. See Refs. [155169] for
further discussion on this topic.
Eq. (1.18.7) is also valid for the case of lossless negative-index media and implies that
the group velocity, and hence the group refractive index ng = c0 /vg , will be positive,
even though the refractive index n is negative. Writing  = || and = || in this



case and noting that = ||/|| and n = ||/ 0 0 , and k = n/c0 , we have:
1

ven






|| d()
|| d(||)
|| d(||)
|| d()
1
+
+
=
|| d
|| d
2
||
d
||
d


d ||
1 d(n)
dk
=
=
=
= vg1
d
c0 d
d
1
=
2

32

1. Maxwells Equations

1.2 Prove the vector analysis identities:

)= 0
(
)= 2 +
(

)= 2 2
(
)A + A
(A)= (
)A + A
(A)= (
A)= 0
(
A)A (
B)
A B = B (
A)2 A
(
A)= (

(Greens rst identity)


(Greens second identity)

1.3 Consider the innitesimal volume element xyz shown below, such that its upper half
=
z.
lies in medium 1 and its lower half in medium 2 . The axes are oriented such that n
Applying the integrated form of Amp`
eres law to the innitesimal face abcd, show that

H2y H1y = Jx z +

Dx
z
t

In the limit z 0, the second term in the right-hand side may be assumed to go to zero,
whereas the rst term will be non-zero and may be set equal to a surface current density,
that is, Jsx limz0 (Jx z). Show that this leads to the boundary condition H1y H2y =
Jsx . Similarly, show that H1x H2x = Jsy , and that these two boundary conditions can be
combined vectorially into Eq. (1.5.4).



from which we also obtain the usual relationship ng = d(n)/d. The positivity of
vg and ng follows from the positivity of the derivatives d()/d and d()/d, as
required to keep ven positive in negative-index media [378].

1.4 Show that the time average of the product of two harmonic quantities A(t)= Re Aejt


and B(t)= Re Bejt with phasors A, B is given by:

A(t)B(t) =

T
0

A(t)B(t) dt =


1
Re AB ]
2

where T = 2/ is one period. Then, show that the time-averaged values of the cross


and dot products of two time-harmonic vector quantities A (t)= Re A ejt and B (t)=
jt
Re B e
can be expressed in terms of the corresponding phasors as follows:

1.19 Problems
1.1 Prove the vector algebra identities:
A (B C)= B(A C)C(A B)
A (B C)= B (C A)= C (A B)
|A B|2 + |A B|2 = |A|2 |B|2
An
+ (n
A) n

A=n

Next, apply the integrated form of Gausss law to the same volume element and show the
boundary condition: D1z D2z = s = limz0 (z).

(BAC-CAB identity)

(
n is any unit vector)

is taken to mean n
(A n
)
An
In the last identity, does it a make a difference whether n
?
A)n
or (n

B(t) =
A (t)B



1
Re A B ,
2

B(t) =
A (t)B



1
Re A B
2

1.5 Assuming that B = H, show that Maxwells equations (1.9.2) imply the following complexvalued version of Poyntings theorem:

(E H )= jH H E Jtot
,

where Jtot = J + jD

1.19. Problems

33

Extracting the real-parts of both sides and integrating over a volume V bounded by a closed
surface S, show the time-averaged form of energy conservation:

1
Re[E H ]dS =
2

] dV
Re[E Jtot
2

which states that the net time-averaged power owing into a volume is dissipated into heat.
For a lossless dielectric, show that the above integrals are zero and provide an interpretation.
1.6 Assuming that D = E and B = H, show that Maxwells equations (1.1.1) imply the following
relationships:




B 
1
E2
Ex + D
= Ex E x
t x
2

where the subscript x means the x-component. From these, derive the following relationship
that represents momentum conservation:

Gx
= Tx
t

1.10 Conductors and plasmas exhibit anisotropic and birefringent behavior when they are in the
presence of an external magnetic eld. The equation of motion of conduction electrons in
v = e(E + v B)mv, with the collisional term
a constant external magnetic eld is m
Ex + y
Ey
z B, and that E = x
included. Assume the magnetic eld is in the z-direction, B =
vx + y
vy .
and v = x
a. Show that in component form, the above equations of motion read:

e
Ex + B vy vx
m
e
y =
Ey B vx vy
v
m
x =
v

where

1
(E2 + H2 )
2

1.7 Show that the causal and stable time-domain dielectric response corresponding to Eq. (1.11.5)
is given as follows:

c. Dene the induced currents as J = Nev. Show that:

where 0 =

2p
0

0 t)u(t)
et/2 sin(

(1.19.2)

0 = 20 2 /4, and we must assume that


where u(t) is the unit-step function and
< 20 , as is typically the case in practice. Discuss the solution for the case /2 > 0 .
1.8 Show that
the plasma frequency for electrons can be expressed in the simple numerical form:
fp = 9 N, where fp is in Hz and N is the electron density in electrons/m3 . What is fp for
the ionosphere if N = 1012 ? [Ans. 9 MHz.]
1.9 Show that the relaxation equation (1.13.2) can be written in the following form in terms of
the dc-conductivity dened by Eq. (1.12.4):
1

r, t)+(
r, t)+
(

(r, t)= 0
0

Then, show that it reduces to the naive relaxation equation (1.7.3) in the limit = 1/ 0.
Show also that in this limit, Ohms law (1.13.1) takes the instantaneous form J = E, from
which the naive relaxation constant relax = 0 / was derived.

where ()=

0
+ j( B )

Ne2
is the dc value of the conductivity.
m

d. Show that the t-domain version of part (c) is:

t
Jx (t)jJy (t)=
0

(t)=

eB
= (cyclotron frequency)
m

e
(Ex jEy )
m
vx jvy =
+ j( B )

Jx jJy = ()(Ex jEy ),

Write similar equations of the y, z components. The quantity Gx is interpreted as the eld
momentum (in the x-direction) per unit volume, that is, the momentum density.

(t)= 0 (t)+0 (t) ,

B =

b. To solve this system, work with the combinations vx jvy . Assuming harmonic timedependence, show that the solution is:

(1.19.1)

where fx , Gx are the x-components of the vectors f = E + J B and G = D B, and Tx is


):
dened to be the vector (equal to Maxwells stress tensor acting on the unit vector x

Tx = Ex E + Hx H x

1. Maxwells Equations

What is the cyclotron frequency in Hz for electrons in the Earths magnetic eld B =
0.4 gauss = 0.4104 Tesla? [Ans. 1.12 MHz.]

D



1
H2
(J B)x +
B x = Hx H x
t
2

fx +

34



(t t ) Ex (t )jEy (t ) dt

where (t)= 0 et ejB t u(t) is the inverse Fourier transform of () and


u(t) is the unit-step function.
e. Rewrite part (d) in component form:

t
Jx (t) =

t
Jy (t) =



xx (t t )Ex (t )+xy (t t )Ey (t ) dt


yx (t t )Ex (t )+yy (t t )Ey (t ) dt

and identify the quantities xx (t), xy (t), yx (t), yy (t).


f. Evaluate part (e) in the special case Ex (t)= Ex u(t) and Ey (t)= Ey u(t), where Ex , Ey
are constants, and show that after a long time the steady-state version of part (e) will
be:

Jx = 0

Ex + bEy
1 + b2

Jy = 0

Ey bEx
1 + b2

1.19. Problems

35

where b = B /. If the conductor has nite extent in the y-direction, as shown above,
then no steady current can ow in this direction, Jy = 0. This implies that if an electric
eld is applied in the x-direction, an electric eld will develop across the y-ends of the
conductor, Ey = bEx . The conduction charges will tend to accumulate either on the
right or the left side of the conductor, depending on the sign of b, which depends on
the sign of the electric charge e. This is the Hall effect and is used to determine the
sign of the conduction charges in semiconductors, e.g., positive holes for p-type, or
negative electrons for n-type.
What is the numerical value of b for electrons in copper if B is 1 gauss? [Ans. 4.3107 .]
g. For a collisionless plasma ( = 0), show that its dielectric behavior is determined from
Dx jDy =  ()(Ex jEy ), where


 ()= 0 1

2p

1.11 This problem deals with various properties of the Kramers-Kronig dispersion relations for
the electric susceptibility, given by Eq. (1.17.8).
a. Using the symmetry properties r ()= r () and i ()= i (), show that
(1.17.8) can be written in the folded form of Eq. (1.17.9).

d
=0
2 2

Hint : You may use the following indenite integral:

(1.19.3)



a + x
dx
1
.

=
ln

a2 x2
2a
a x

c. Using Eq. (1.19.3), show that the relations (1.17.9) may be rewritten as ordinary integrals (without the P instruction) as follows:

r () =

i () =

 i ( )i ()
d
2 2

r ( )r ()
d
2 2

(1.19.4)

Hint : You will need to argue that the integrands have no singularity at  = .
d. For a simple oscillator model of dielectric polarization, the susceptibility is given by:

() = r ()ji ()=
=

2p
20 2 + j

2p (20 2 )
(20 2 )2 +2 2

2p

(1.19.5)

(20 2 )2 +2 2

Show that for this model the quantities r () and i () satisfy the modied KramersKronig relationships (1.19.4). Hint : You may use the following denite integrals, for
which you may assume that 0 < < 20 :
2

(20

dx
1
=
,
x2 )2 +2 x2
20

x2 dx
1
=

(20 x2 )2 +2 x2

Indeed, show that these integrals may be reduced to the following ones, which can be
found in standard tables of integrals:
2

dy
1

2y2

cos +

y4

y2 dy
1
= 
cos + y4
2(1 cos )

2y 2

where sin(/2)= /(20 ).


e. Consider the limit of Eq. (1.19.5) as 0. Show that in this case the functions r , i
are given as follows, and that they still satisfy the Kramers-Kronig relations:

r ()= P

2p
0

+P

2p
0 +

i ()=

2p
2 0

( 0 )( + 0 )

1.13 An isotropic homogeneous lossless dielectric medium is moving with uniform velocity v with
respect to a xed coordinate frame S. In the frame S moving with dielectric, the constitutive
relations are assumed to be the usual ones, that is, D  = E  and B  = H  . Using the Lorentz
transformations given in Eq. (H.30) of Appendix H, show that the constitutive relations take
the following form in the xed frame S:
D = E + av (H v E) ,

b. Using the denition of principal-value integrals, show the following integral:

1. Maxwells Equations

1.12 Derive the Kramers-Kronig relationship of Eq. (1.17.7) by starting with the causality condition (t)u(t)= 0 and translating it to the frequency domain, that is, expressing it as the
convolution of the Fourier transforms of (t) and u(t).

( B )

where p is the plasma frequency. Thus, the plasma exhibits birefringence.

36

B = H av (E + v H) ,

a=

 0 0
1 v2

Das könnte Ihnen auch gefallen