You are on page 1of 12

Chapter I

Wave Theory of Optical


Waveguides

The basic concepts and equations of electromagnetic wave theory required for
the comprehension of lightwave propagation in optical waveguides are presented.
The light confinement and formation of modes in the waveguide are qualitatively
explained, taking the case of a slab waveguide. Maxwell's equations, boundary
conditions, and the complex Poynting vector are described as they form the basis
for the following chapters.

1.1. WAVEGUIDE STRUCTURE

Optical fibers and optical waveguides consist of a core, in which light is


confined, and a cladding, or substrate surrounding the core, as shown in Fig. 1.1.
The refractive index of the core n~ is higher than that of the cladding n 0.
Therefore the light beam that is coupled to the end face of the waveguide is
confined in the core by total internal reflection. The condition for total internal
reflection at the core-cladding interface is given by n 1 sin('rr/2- ~) ~>n 0. Since
the angle ~ is related with the incident angle 0 by sin 0 = n 1 sin ~ ~< v/n12 - n~,
we obtain the critical condition for the total internal reflection as

0 ~< sin -1 ~/nl2 - n~ ~ Omax. (1.1)

The refractive-index difference between core and cladding is of the order of


n 1 - n o = 0.01. Then 0max in Eq. (1.1) can be approximated by

Omax~ ~/n12 -- n 2. (1.2)


2 Wave Theory of Optical Waveguides

no

~=Cor'e nI nl
.~ x=O
y z z i'1
x=-a
Cladding no no

Figure 1.1 Basic structure and refractive-index profile of the optical waveguide.

x denotes the maximum light acceptance angle of the waveguide and is known
0ma
as the numerical aperture (NA).
The relative refractive-index difference between n I and n o is defined as

A -- n2 - n2 = ~n l -. no (1.3)
2n 2 n1

A is commonly expressed as a percentage. The numerical aperture NA is related


to the relative refractive-index difference A by

NA-- 0max '~ n l ~ . (1.4)

The maximum angle for the propagating light within the core is given by
+max ~ 0max/nl ~ x / ~ . For typical optical waveguides, NA = 0.21 and 0max =
12~ 8.1 ~ when n I = 1.47, A = 1% (for n o = 1.455).

1.2. F O R M A T I O N OF GUIDED MODES

We have accounted for the mechanism of mode confinement and have indi-
cated that the angle + must not exceed the critical angle. Even though the angle
is smaller than the critical angle, light rays with arbitrary angles are not
able to propagate in the waveguide. Each mode is associated with light rays
at a discrete angle of propagation, as given by electromagnetic wave analysis.
Here we describe the formation of modes with the ray picture in the slab wave-
guide [1 ], as shown in Fig. 1.2. Let us consider a plane wave propagating along
the z-direction with inclination angle +. The phase fronts of the plane waves are
perpendicular to the light rays. The wavelength and the wavenumber of light in
the core are h / n 1 and k n l ( k = 2~r/h), respectively, where h is the wavelength
of light in vacuum. The propagation constants along z and x (lateral direction)
are expressed by

-- kn 1 c o s ~), (1.5)

K = kn 1 sin +. (1.6)
Formation of Guided Modes 3

x Light ray
Phase front \
y*--- z R Q/ \ no

x t / ' l -~ \. , v " \ \ \/~P "~ k',\ \


', \ ',. ~ . nl 2a
|

S no
Figure 1.2 Light rays and their phase fronts in the waveguide.

Before describing the formation of modes in detail, we must explain the


phase shift of a light ray that suffers total reflection. The reflection coefficient
of the totally reflected light, which is polarized perpendicular to the incident
plane (plane formed by the incident and reflected rays), as shown in Fig. 1.3, is
given by [2]

Ar n 1 sin + + jv/n~ cos 2 + - n~


r. . . . (1.7)
Ai n 1 sin + -- jv/n~ cos 2 + -- no2

When we express the complex reflection coefficient r as r - e x p ( - j ~ ) , the


amount of phase shift 9 is obtained as

-- - 2 t a n -1 v/n2 COS2 + - - n~ --2tan -1 1 (1.8)


n 1 sin + sin-5 ~ "

where Eq. (1.3) has been used. The foregoing phase shift for the totally reflected
light is called the Goos-H~inchen shift [1, 3].
Let us consider the phase difference between the two light rays belonging to
the same plane wave in Fig. 1.2. Light ray PQ, which propagates from point P
to Q, does not suffer the influence of reflection. On the other hand, light ray RS,

~Z

H~ H
Incident light A~ Reflected light A r
Figure 1.3 Total reflection of a plane wave at a dielectric interface.
4 Wave Theory of Optical Waveguides

propagating from point R to S, is reflected two times (at the upper and lower
core-cladding interfaces). Since points P and R or points Q and S are on the
same phase front, optical paths PQ and RS (including the Goos-H~inchen shifts
caused by the two total reflections) should be equal, or their difference should
be an integral multiple of 2"rr. Since the distance between points Q and R is
2a/tan + - 2a tan +, the distance between points P and Q is expressed by

(2a ) ( 1 )
el "- tan ~ - 2a tan ~ cos ~ = 2a sin ~ - 2 sin ~ . (1.9)

Also, the distance between points R and S is given by

2a
e 2 = q~sin---
v. (1.1 O)

The phase-matching condition for the optical paths PQ and RS then becomes

(kn Ie 2 + 20) - k n 1e I = 2m'rr, (1.11)

where m is an integer. Substituting Eqs. (1.8)-(1.10) into Eq. (1.11) we obtain


the condition for the propagation angle ~ as

m'rr ~ 2A 1. (1.12)
tan(knlasind?---~-)- sin2tb

Equation (1.12) shows that the propagation angle of a light ray is discrete
and is determined by the waveguide structure (core radius a, refractive index
n 1, refractive-index difference A) and the wavelength k of the light source
(wavenumber is k = 2"rr/k) [4]. The optical field distribution that satisfies the
phase-matching condition of Eq. (1.12) is called the mode. The allowed value of
propagation constant [3 [Eq. (1.5)] is also discrete and is denoted as an eigenvalue.
The mode that has the minimum angle ~ in Eq. (1.12) (m =0) is the fundamental
mode; the other modes, having larger angles, are higher-order modes (m ~> 1).
Figure 1.4 schematically shows the formation of modes (standing waves) for
(a) the fundamental mode and (b) a higher-order mode, respectively, through
the interference of light waves. In the figure the solid line represents a positive
phase front and a dotted line represents a negative phase front, respectively. The
electric field amplitude becomes the maximum (minimum) at the point where two
positive (negative) phase fronts interfere. In contrast, the electric field amplitude
becomes almost zero near the core-cladding interface, since positive and negative
phase fronts cancel out each other. Therefore the field distribution along the
x-(transverse) direction becomes a standing wave and varies periodically along
the z direction with the period hp--()k/nl)/COSdp --27r/f3.
Formation of Guided Modes 5

, -,~ Phase front


Cladding no ~ ~ / ~ /
Core nl /~ ?\ /~ /\ il~ /~, /1~ A, 11~? \
kn, / '~/ '~i y V V V V VV\ E
x~ \ /\ /~ /\ / ~ A A A /~,A/
'rr- o V V V V ' T V V V ' T V
L _l ~o~
E I- -1 2 =2~r/#o

(a) Fundamental mode ( m = O )

~ ~ Phase front
x
Cladding
Core
knl
~\\./"~",,V/'N~//Xx</'V',,,/ b

// \

h Jq 2(p1)=2~/~1
E I-

(b) Higher-order mode (m = 1 )


Figure 1.4 Formationof modes: (a) Fundamental mode, (b) higher-order mode.

Since nl sin + = sin 0 ~<v/n12 - no2 from Fig. 1.1, Eqs. (1.1) and (1.3) give the
propagation angle as sin ~ ~< C2-A. When we introduce the parameter

sin +
~-~, (1.13)

which is normalized to 1, the phase-matching Eq. (1.12) can be rewritten as

knla~/~-A_ cos -1 ~ + m'rr/2


. (1.14)

The term on the left-hand side of Eq. (1.14) is known as the normalized
frequency, and it is expressed by

v=knla~-A. (1.15)

When we use the normalized frequency v, the propagation characteristics of the


waveguides can be treated generally (independent of each waveguide structure).
6 Wave Theory of Optical Waveguides

The relationship between normalized frequency v and ~ (propagation constant [3),


Eq. (1.14), is called the dispersion equation. Figure 1.5 shows the dispersion
curves of a slab waveguide. The crossing point between Xl= (cos -1 ~ + m~r/2)/~
and ~q--v gives ~mfor each mode number m, and the propagation constant 13m
is obtained from Eqs. (1.5) and (1.13).
It is known from Fig. 1.5 that only the fundamental mode with m - 0 can
exist when v < vC= "rr/2. v c determines the single-mode condition of the slab
waveguide--in other words, the condition in which higher-order modes are cut
off. Therefore it is called the cutoff v-value. When we rewrite the cutoff condition
in terms of the wavelength we obtain

2"rr
kc= ~anl~2-~. (1.16)
Vc

10

m=3
8

m= 2

m- 1

0 v m 4.

m--O

00 .2 .4. .6 .8 1

Figure 1.5 Dispersion curves of a slab waveguide.


Maxwell's Equations 7

h c is called the cutoff(free-space) wavelength. The waveguide operates in a single


mode for wavelengths longer than hc. For example, h c - - 0 . 8 Ixm when the core
width 2 a - 3.54 Ixm for the slab waveguide of n 1 - 1.46, A = 0.3%(n 0 = 1.455).

1.3. MAXWELL'S EQUATIONS

Maxwell's equations in a homogeneous and lossless dielectric medium are


written in terms of the electric field e and magnetic field h as [5]

Oh
V x e - - I x Ot' (1.17)

0e
V x h -- e Ot' (1.18)

where e and tx denote the permittivity and permeability of the medium, respec-
tively, e and Ix are related to their respective values in a vacuum of e 0 -
8.854 • 10-~2[F/m] and I x 0 - 47r • 10-7[H/m] by

8 - - 80 n2, (1.19a)

IX- Ixo, (1.19b)

where n is the refractive index. The wavenumber of light in the medium is then
expressed as [5]

F= co~= con e~/-~o~o- kn. (1.20)

In Eq. (1.20), co is an angular frequency of the sinusoidally varying electromag-


netic fields with respect to time; k is the wavenumber in a vacuum, which is
related to the angular frequency co by
CO
k = co~/eoix0 = --. (1.21)
C

In Eq. (1.21), c is the light velocity in a vacuum, given by

1
c- ~ = 2.998 x lOS[m/s]. (1.22)
~/~:o~o
The fact that the units for light velocity c are m/s is confirmed from the units of
the permittivity e 0 [F/m] and permeability Ix0 [H/m] as

1 m m m

v/[F/m][H/m] v/-F 9H v/[A. s/V][V, s/A] s


8 Wave Theory of Optical Waveguides

When the frequency of the electromagnetic wave is f[Hz], it propagates


c/f [m] in one period of sinusoidal variation. Then the wavelength of electro-
magnetic wave is obtained by

c to/k 2'rr
h= f f k ' (1.23)

where to = 2"rrf.
When the electromagnetic fields e and h are sinusoidal functions of time,
they are usually represented by complex amplitudes, i.e., the so-called phasors.
As an example consider the electric field vector

e(t) = [El cos(cot + 6), (1.24)

where IEI is the amplitude and + is the phase. Defining the complex amplitude
of e(t) by

E - IEId+, (1.25)

Eq. (1.24) can be written as

e(t) = Re{EeJ't}. (1.26)

We will often represent e(t) by

e( t) = Ee -i''t (1.27)

instead of by Eq. (1.24) or (1.26). This expression is not strictly correct, so


when we use this phasor expression we should keep in mind that what is meant
by Eq. (1.27) is the real part of Ee j~ In most mathematical manipulations,
such as addition, subtraction, differentiation and integration, the replacement of
Eq. (1.26) by the complex form (1.27) poses no problems. However, we should
be careful in the manipulations that involve the product of sinusoidal functions.
In these cases we must use the real form of the function (1.24) or complex
conjugates [see Eqs. (1.42)].
When we consider an electromagnetic wave having angular frequency to and
propagating in the z direction with propagation constant [3, the electric and
magnetic fields can be expressed as

e = E ( r ) d (''t-f~z), (1.28)
h = H ( r ) e j(~ . (1.29)
Maxwell's Equations 9

where r denotes the position in the plane transverse to the z-axis. Substituting
Eqs. (1.28) and (1.29) into Eqs. (1.17) and (1.18), the following set of equations
are obtained in Cartesian coordinates:

OEz
-~y + jf3Ey = - j o , ~ o H x

OEz
-jf3G- o--;=-jo~ony
Oey Oex
= -flOP~oHz
Ox Oy
(1.30)
OHz
-~y + j[3Hy = flOeonZEx

-J~3Hx - OHz
0---~- jtoso nz Ey

Ony Onx = jt08o n2 E z .


Ox Oy

The foregoing equations are the bases for the analysis of slab and rectangular
waveguides.
For the analysis of wave propagation in optical fibers, which are axially
symmetric, Maxwell's equations are written in terms of cylindrical coordinates:

10E z
r O0 + j~3E~ = -J~176

_j~3E r OEz_ jtolxoHo


Or
1 0 10E r
r ~(rEo) r 80 = - f l ~ 1 7 6
(1.31)
1 OHz
r O0 + jf3H~ = fl~176

-jf3Hr- OHz
0-7 =flOe, on2E o
1 0 IOH r
r -dr (rH~ - -r ~O0 = jo,~o n~ Ez.

Maxwell's Eqs. (1.30) or (1.31) do not determine the electromagnetic field


completely. Out of the infinite possibilities of solutions of Maxwell's equations,
we must select those that also satisfy the boundary conditions of the respective
problem. The most common type of boundary condition occurs when there are
discontinuities in the dielectric constant (refractive index), as shown in Fig. 1.1.
10 Wave Theory of Optical Waveguides

At the boundary the tangential components of the electric field and magnetic
field should satisfy the conditions

E} 1) --g} 2) (1.32)
H~ 1) --H} 2) , (1.33)

where the subscript t denotes the tangential components to the boundary and
the superscripts (1) and (2) indicate the medium, respectively. Equations (1.32)
and (1.33) mean that the tangential components of the electromagnetic fields
must be continuous at the boundary. There are also natural boundary conditions
that require the electromagnetic fields to be zero at infinity.

1.4. P R O P A G A T I N G POWER

Consider Gauss's theorem (see Section 10.1) for vector A in an arbitrary


volume V

f f v . A dv -- f f A. n ds, (1.34)
v S

where n is the outward-directed unit vector normal to the surface S enclosing


V and dv and ds are the differential volume and surface elements, respectively.
When we set A - - e x h in Eq. (1.34) and use the vector identity

V.(e x h) = h . V x e - e . V xh, (1.35)

we obtain the following equation for electromagnetic fields:

ff f (h . V • e - e . V x h)dv = f f
v S
(e x h ) . n ds. (1.36)

Substituting Eqs. (1.17) and (1.18) into Eq. (1.36) results in

f f f ( ~e . -~
8e + p~h . -~
8 h ) d r - - f f (e x h ) . n ds. (1.37)
V S

The first term in Eq. (1.37)

8e 8 (~ ) SWe
(1.38)
~e . Ot : Ot -~e . e = Ot '
Propagating Power 11

represents the rate of increase of the electric stored energy W e and the second term

t x h . Ot = Ot -~h.h = --'Ot (1.39)


represents the rate of increase of the magnetic stored energy W h, respectively.
Therefore, the left-hand side of Eq. (1.37) gives the rate of increase of the
electromagnetic stored energy in the whole volume V; in other words, it repre-
sents the total power flow into the volume bounded by S. When we replace the
outward-directed unit vector n by the inward-directed unit vector U z ( = - n ) , the
total power flowing into the volume through surface S is expressed by

P=ff -(ex h).nds--ff ( e x h ) . u zds. (1.40)


S S

Equation (1.40) means that e x h is the vector representing the power flow,
and its normal component to the surface (e x h) "Uz gives the amount of power
flowing through unit surface area. Therefore, vector e x h represents the power-
flow density, and

S = e • h[W/m 2] (1.41)

is called the P o y n t i n g vector. In this equation, e and h denote instantaneous


fields as functions of time t. Let us obtain the average power-flow density in an
alternating field. The complex electric and magnetic fields can be expressed by

1
e ( t ) - R e { E e j''t } -- -~{EeJ't + E* e -J~" } , (1.42a)

h(t)- Re{He TM}-- ~1 {HeJ~ + H* e -jt~ }, (1.42b)

where * denotes the complex conjugate. The time average of the normal com-
ponent of the Poynting vector is then obtained as

(S. Uz) = ((e x h). Uz)


1
= -~ ([(EeJC~ E*e-J ~ • (HeJt~ H*e-Jtot)]. Uz}

1 1
= ~(E • H* + E* x H ) . u z = ~ Re{(E x H*). Uz}, (1.43)

where () denotes a time average. Then the time average of the power flow is
given by

P = f f ~Re{(E x H*). Uz} ds. (1.44)


S
12 Wave Theory of Optical Waveguides

Since E x H* often becomes real in the analysis of optical waveguides, the time
average propagation power in Eq. (1.44) is expressed by

P = f f l ~ (E • I-I*). u z ds. (1.45)


s

REFERENCES

[1] Marcuse, D. 1974. Theory of Dielectric Optical Waveguides. New York: Academic Press.
[2] Born, M. and E. Wolf. 1970. Principles of Optics. Oxford: Pergamon Press.
[3] Tamir, T. 1975. Integrated Optics. Berlin: Springer-Verlag.
[4] Marcuse, D. 1972. Light Transmission Optics. New York: Van Nostrand Rein-hold.
[5] Stratton, J. A. 1941. Electromagnetic Theory. New York: McGraw-Hill.

You might also like