Theory of Optical Waveguides: With 32 Figures
Theory of Optical Waveguides: With 32 Figures
Theory of Optical Waveguides: With 32 Figures
H. Kogelnik
With 32 Figures
The simplest dielectric guide is the planar slab guide shown in Fig. 2.1,
where a planar film of refractive index nf is sandwiched between a substrate
and a cover material with lower refractive indices ns and nc (nf> ns ~ nc).
Often the cover material is air, in which case nc = 1. As an illustration, we
have listed in Table 2.1 the refractive indices of some dielectric waveguide
materials used in integrated optics. Typical differences between the indices
of the film and the substrate range from 10- 3 to 10-1, and a typical film
thickness is 1 /Lm. The light is confined by total internal reflection at the
film-substrate and film-cover interfaces.
Dielectric waveguides have already been the subject of several text-
books, [2.1-9], and we can refer the reader to these for a history on the
subject as well as for a more complete list of references.
7
T. Tamir (ed.), Guided-Wave Optoelectronics
© Springer-Verlag Berlin Heidelberg 1988
Table 2.L Refractive index n of optical waveguide materials
Dielectric material
8
2.1 Ray Optics of the Slab Waveguide
In this section, we propose to discuss and develop the ray-optical model of
light propagation in a slab waveguide. Ray-optical techniques in connection
with slab waveguides have been explored and used by Tien [2.17], Maurer
and Felsen [2.18]' Lotsch [2.19] and others. We have chosen the slab wave-
guide, sketched in Fig. 2.1, for two reasons: first, because it is relatively
easy to understand and analyze, and second because it is of considerable
practical interest in integrated optics. We shall use the ray-optical picture to
introduce the basic concepts and terminology of dielectric-waveguide theory,
including the nature of the modes of propagation, waveguide cutoff, the
propagation constants and the effective guide thickness. In addition, we
employ this picture to derive and provide a number of results of interest
to the experimenter, such as plots of the propagation constant and of the
effective width of slab waveguides. The ray-optical picture is a very simple
picture with great intuitive appeal, but it is not as complete a description as
that provided by electromagnetic theory, which we discuss later in Sects. 2.2
and 3. However, the results we present here are in perfect agreement with
the latter.
Our picture of light guidance in a slab waveguide is one of light rays
tracing a zig-zag path in the film, with total internal reflection of the light
occurring at the film-substrate and film-cover interfaces. As reflection and
refraction at these dielectric interfaces play an important role in the guiding
process, let us briefly review the relevant laws and their consequences.
9
interface, is partially reflected and refracted as shown. The exit angle ()2 of
the refracted wave C is given by Snell's law
(2.1.1)
The reflected wave has a complex amplitude B at the interface linearly
related to A by a complex reflection coefficient R
B =R·A (2.1.2)
The reflection coefficient depends on the angle of incidence and the polariza-
tion of the light, and is given by the Fresnel formulas. For TE polarization
(i.e., electric fields perpendicular to the plane of incidence spanned by the
wave normal and the normal to the interface) we have
J
ni sin 2 (Jl - n~
tan<pTE = - ' - - - - - - - (2.1.7)
nl cos (Jl
2 . / n2 sin 2 (Jl - n2
tan <PTM = -nl .V2 -1- - - - - - , - - -2 (2.1.8)
n~ nl cos (Jl
Figure 2.3 shows the dependence of <PTE on the angle of incidence (Jl for a
selection of index ratios n2/nl where the values 0.3, 0.5 and 0.7 correspond
10
2"-r----------------:a
7r
f
CPTE -1
al radiation mode
OL--L-L2~O-O~-4~O~o~-J60-0~-~~
b I substrate mode
Fig. 2.3. Phase shift <PTE of the TE mode as
a function of the angle of incidence 01
approximately to interfaces between air and GaAs, LiNb03, and Si02, re-
spectively. We note that the phase shift increases from 0 at the critical angle
to 7r /2 at grazing incidence (lh = 90 0 ). It increases with infinite slope at
B1 = Be and a slope of (1- nVnr)-1/2 at B1 = 90 0 • The behaviour of cPTM
is quite similar.
Consider, now, the ("asymmetric") slab waveguide structure shown in
Fig. 2.4 with a film of index nf and substrate and cover materials of index
ns and nco In general, we have nf> ns > nc and two critical angles of in-
terest, Bs for total reflection from the film-substrate interface and Be < Bs
for total reflection from the film-cover interface. When we examine what
happens as the angle of incidence B is increased, we discover that there are
three distinct cases which are sketched in Fig. 2.4. For small angles B < Bs , Be
light incident from the substrate side is refracted according to Snell's law
and escapes from the guide through the cover (a). There is essentially no
confinement of light, and the electromagnetic mode corresponding to this
picture is called a "radiation mode" which is discussed in more detail in
Sect. 2.3. When B is increased somewhat, such that Be < B < Bs , we then find
the situation depicted in (b). The light incident from the substrate is re-
fracted at the film-substrate interface, totally reflected at the film-cover
interface, refracted back into the substrate through which the light escapes
from the structure. Again, there is no confinement and we talk of a "sub-
strate radiation mode" (Sect. 2.3). Finally (c), when B is large enough, we
have Bs , Be < B, i.e., total internal reflection at both interfaces. The light,
11
Fig. 2.5. Side-view of a slab
x COVER
waveguide showing wave nor-
mals of the zig-zag waves cor-
responding to a guided mode
z
SUBSTRATE
For a guided mode of the slab guide, the zig-zag picture predicts a propa-
gation constant (J (and the related phase velocity v p )
(J = w/vp = knfsin () (2.1.11)
which is the z-component of the wave vector knf. However, not all angles
() are allowed; only a discrete set of angles (and sometimes none) lead to a
self-consistent picture that corresponds to what we call the "guided modes" .
To examine this in more detail, let us look at a guide cross-section z = const
and add up the phase shifts that occur as we move up from the lower film
boundary (x = 0) with one wave to the other boundary (x = h) and then
12
back down again with the reflected wave to where we started from. For
self-consistency, the sum of all these phase shifts must be a multiple of 27r.
For a film of thickness h we have, specifically, a phase shift of knfh cos ()
for the first transverse passage through the film, a phase shift of -2rPc
on total reflection from the film-cover interface, another knfh cos () on the
transverse passage down, and a phase shift of - 2rPs on total reflection from
the film-substrate boundary. Thus, we have the self-consistency condition
(also known as the "transverse resonance condition")
2knfh cos () - 2rPs - 2rPc = 2V7r (2.1.12)
where v is an integer (0,1,2 ... ) which identifies the mode number. As dis-
cussed before, the phase shifts rPs and rPc are functions of the angle 8, as
described by (2.1.7 and 8) after the appropriate substitutions for nl and
n2. The above equation is essentially the dispersion equation of the guide
yielding the propagation constant f3 as a function of frequency wand film
thickness h. From (2.1.5 and 11), we find for guided modes that f3 is bounded
by the plane-wave propagation constants of substrate and film
kns < f3 < knf . (2.1.13)
which is bounded by
ns <N <nf . (2.1.15)
Figure 2.6 sketches a graphical solution of the dispersion equation (2.1.12)
for the fundamental mode (v = 0) which gives us further information of the
propagation characteristics of the guide. We have drawn here, as a function
of the angle 8, both the phase shift on film traversal knfh cos 8 (dotted
8-
O· 20· 4 O' 60' 90'
r -____L-____L -____L-____- - , ~
13
curve) and the sum of the phase shifts (4)s + 4>c) at the film boundaries.
We show the latter for two cases, the symmetric guide where 4>c = 4>s (solid
curve), and the asymmetric guide (dashed curve). Consider the symmetric
guide first, where the intersection between the solid and the dotted curve
yields the zig-zag angle 0 of the fundamental mode. We note that the zig-zags
get steeper (0 smaller) as hi>' gets smaller, but there is always a solution
even when the film thickness gets very small. This implies that there is
no cutoff for the fundamental mode of a symmetric guide. Of course, as the
guide gets thicker, it supports more and more guided modes. Considering the
asymmetric guide, we look for an intersection between the dotted and the
dashed curve. However, only the portion of the (4)s + 4>c) curve emphasized
by shading is above the critical angle Os of the film-substrate interface. For
sufficiently thin films, we do not get an intersection of the curves above
cut-off, which implies that an asymmetric guide cannot always support a
guided mode, i.e., there is a cut-off condition even for the fundamental.
FORBIDDEN
REGION
Figure 2.7 shows a sketch of an w-{3 diagram that restates some of the
above-discussed dispersion characteristics, which are typical for a dielectric
waveguide. The first three guided modes (v = 0,1,2) are shown. At the
cut-off frequency, the propagation constants assume the value of the lower
bound nsk, and as w (or the thickness h) increases, {3 approaches its upper
bound nfk and more and more guided modes exist. In addition to the dis-
crete spectrum of the guided modes, the diagram also shows the continuous
spectrum of the radiation modes.
To obtain a more precise w-{3 diagram fo the asymmetric slab guide, we
have to evaluate (2.1.12) numerically. To make the results of such a numer-
ical evaluation more broadly applicable, we introduce normalizations that
combine several guide parameters. First, we define a normalized frequency
and film thickness V by
(2.1.16)
14
and then a normalized guide index b related to the effective index N (and
fi) by
(2.1.17)
The index b is zero at cut-off and approaches unity far away from it. For
small index differences (nf - ns) we have the simple linear relation
(2.1.18)
Table 2.2. Asymmetry measures for the TE modes (aE) and the TM modes (aM) of
slab waveguides
Waveguide ns nf nc aE aM
For the TE modes, we use (2.1.7) together with the above normaliza-
tions to write the dispersion relation (2.1.12) in the form
vVl=b = V'Tr + tan- 1 Jb/(1- b) + tan-1J(b + a)/(l- b) .(2.1.20)
15
v=o
.4 r--t-/ffl-+
II I r; jI V/, V
/1--+---+------11+-/,1/-H-/-+--+--II--+----1
/I~---+----'/'+f-!.+tf
I /I
.3 t-----i-t--+t
1/ /11/ IIIX/
/+-+--tf--/HJi/'+-+-+--;/Ih',;;'+,/Y---I-+-+-+---+---I
/1/++'
II II / II I. IIIj-+-+--l------l-+-+--I
/1
.2 t--------I'--t/-t-I/f--+---t---.jf-++j/++---+--/"'/'-1-1-
II = 2: In; - n; (2.l.24)
For the TM mode, we get cut-off conditions of the same form as for
the TE mode and w-{3 diagrams that are very similar. In fact, when the
index differences (nf - ns) are small, we can apply the diagram of Fig. 2.8
to the TM modes. However, these statements are only correct if we define
the asymmetry measure in a somewhat different manner [2.20], namely, by
a--L.
n4 ns2 _ nc2
- n4c nf2 _ ns2 (2.l.25)
Ilustrative values for this are given in Table 2.2 under aM.
16
2.1.3 The Goos-Hanchen Shift
So far, we have described the light in the waveguide in terms of plane waves
and their wave normals and phases. In this subsection and in the next, we
consider also the energy of the light and its flow through the guide. To
prepare for this, we have to be more precise about what we mean by a light
ray. A light ray is defined here as the direction of the Poynting vector or the
energy flow of light. Consistent with this is the view of a ray as the axis of
a narrow beam of light or wave packet. The relation between wave normal
and ray is essentially the spatial analog of the relation between the phase
velocity and the group velocity. For the simple case of a plane wave in a
homogeneous, isotropic medium the directions of the wave normal and the
ray are the same, but in an anisotropic medium the ray generally points in
a direction different from that of the wave normal.
The Goos-Hiinchen shift that occurs on total reflection from a dielectric
interface is another case where the ray behaves differently than the wave
normal. Here the reflected ray (B) is shifted laterally relative to the incident
ray or wave packet (A), as indicated in Fig. 2.9. This lateral ray shift has
turned out to be an important element in the understanding of the flow of
energy in dielectric waveguides in terms of the ray picture.
To determine the lateral ray shift, shown as 2zs in Fig. 2.9, consider a
simple wave packet consisting of two plane waves incident at two slightly
different angles. If the z-components of the corresponding wave vectors are
(3 ± 11(3, we can write for the complex amplitude A(z) of the incident wave
packet at the interface x = 0
A = [exp(jl1(3z) + exp( -jl1(3z)]exp( -j(3z)
= 2cos(11(3z)exp( -j(3z) (2.1.26)
Before applying the reflection laws (2.1.2 and 6) to each individual plane
wave, we have to remember that the phase shift </> occuring on total reflection
is a function of () (and (3). For smalll1</> and 11(3, we can use an expansion
of the from
17
<fJ({3 +.1(3) = <fJ({3) + ~;.1{3 . (2.1.27)
With this, we obtain for the amplitude B(z) ofthe reflected wave packet at
x=o
B = {exp[j(.1(3z - 2.1<fJ)] + exp[ - j(.1{3z - 2.1<fJ)]}exp[ - j({3z - 2<fJ)]
= cos [.1{3(z - 2zs)]exp[ - j({3z - 2<fJ)] (2.1.28)
where
d<fJ
Zs = d{3 (2.1.29)
This gives us the lateral shift of the wave packet, i.e., of the ray, in compact
and simple form [2.21,22]. Using (2.1.11,7 and 8), we obtain for the TE
modes
kzs = (N 2 - n;)-1/2 tan B , (2.1.30)
and for the TM modes
kzs = (N 2 - ns)
2 -1/2 tan B j(N2 N2n - 1)
""2 + ""2
ns
(2.1.31)
f
As sketched in Fig. 2.9, this lateral ray shift would indicate that the light
penetrates to a depth Xs into the substrate before it is reflected, where
Zs
xs -- - -
tanB (2.1.32)
If we compare this result with the electromagnetic field solutions to be given
in Sect.2.3, we find that these predict evanescent fields in the substrate
whose decay constants are closely related to this penetration depth Xs of
the ray.
18
Fig.2.10. Ray picture of
COVER zig-zag light propagation
in a slab waveguide. Goos-
Hanchen shifts are incor-
porated in the model, and
the effective guide thick-
ness heff is indicated
SUBSTRATE
(2.1.34)
as a function of the normalized frequency V. For TE modes we have
H = V+1/vb+1/~ (2.1.35)
The corresponding plots are shown in Fig. 2.11 for four values of the asym-
-IN
~Ul 7
c:
(\II
r=
-.,
6
..c:
-""
II 5
:r:
"" ""
,/
4 ""
3
2 3 4 5 6
I
V = kh( nf~ n5 2 )2:
Fig. 2.11. Normalized effective thickness of a slab waveguide as a function of the nor-
malized film thickness V for various degrees of asymmetry (after [2.20])
19
metry measure. Similar plots can be obtained for the TM modes [2.20]. In
Fig. 2.11, we note the occurrence of minimum values of H(V), for which
we obtain maximum confinement of the light. For highly asymmetric guides
(a = 00), for example, we have a minimum of Hmin = 4.4 at V = 2.55. This
implies a minimum effective thickness of
(h eff / /\\).
mm -07(
- . nf2 - ns2)-1/2 (2.1.36)
For a typical thin-film glass waveguide, we have ns 1.5, nf = 1.6 and
(heff/ 'x)min ~ 1.3.
\7 X if = aD/at (2.2.2)
where t is time, \7 = (a/ax, a/ay, a/az) is the del operator, and E(t), if(t) ,
D(t) and E(t) are the time-dependent vectors of the dielectric and magnetic
field, the electric displacement and the magnetic induction, respectively. We
assume fields with a periodic time dependence which we write in the form
E(t) = Eexp(jwt) + E*exp( -jwt) etc. (2.2.3)
where E is a complex amplitude, w the angular frequency, and the asterisk
indicates a complex conjugate. Assuming a lossless medium with a scalar
dielectric constant c:(w) and a scalar magnetic permeability p" we have the
constitutive relations
20
D=c:E (2.2.4)
B=f..LH (2.2.5)
With this we get. Maxwell's equations for the complex amplitudes of the
form
\7 X E= -jwf..LH (2.2.6)
\7 X H=jwc:E (2.2.7)
for the fields EI, E2, etc., at the boundary. In dielectric waveguides we
usually have a constant permeability f. L = f..Lo, which implies equality of the
magnetic field vectors HI = H2 at the boundary.
z
€(X,y)
E= Et +Ez
21
~t X Et = -jwp,Hz ~t X Ht = jwc:Et (2.2.10)
where v is a mode label (indicating the mode number, for example), and,Bv
is the propagation constant of the mode. For guides providing confinement
in two dimensions, such as strip guides or fibers, we need, of course, two
labels, but only one is shown here, for simplicity. Inserting the modal fields
of (2.2.15) into Maxwell's equations (2.2.10-12), we obtain
~t X Htll = jwc:Ezv (2.2.16)
(2.2.17)
(2.2.18)
The general nature of the solutions to these equations was discussed
in detail by McKenna [2.24] and in textbooks, [2.1-9]. One encounters a
situation which is analogous to that encountered in quantum mechanics
where one seeks solutions to Schrodinger's equation for various potential
distributions and finds two types of solutions, one corresponding to bound
states and the other to unbound states. For dielectric waveguides, we find
22
guided modes (bound states) where the energy is confined near the axis, and
radiation modes (unbound states) with their energy spread out through the
medium surrounding the guide. The guided modes are associated with a
discrete spectrum of propagation constants fJv, while the radiation modes
belong to a continuum. One also finds evanescent modes with imaginary
propagation constants fJv = -ja v and which decay as exp( -avz). Solutions
for specific waveguide examples are given in Sects. 2.3 and 4.
(2.2.25)
(2.2.26)
23
and a wave equation for H y of the form
which is true for all isotropic dielectric waveguides. By reversing the sign of
z in Maxwell's equations (2.2.10-12), we can construct new solutions of the
form
Et2(Z) = Et1( -z) E Z 2(Z) = -Ezl(-Z) (2.2.31)
Ht2(Z) = -Ht1( -z) Hz2(Z) = Hzl( -z) (2.2.32)
where we have omitted to indicate the (x, y) dependence in order to empha-
size the z-reversal operation.
In applying these reversal operations to the modal fields of the form
of (2.2.15) we have to distinguish between the propagating modes with real
valued f3, and the evanescent modes with imaginary f3. In the first case, a
forward traveling mode will vary as exp( -jf3z) and either time reversal or
z-reversal yield a backward traveling mode varying as exp(jf3z). As the new
solution must be unique, the application of (2.2.29), as well as (2.2.31,32)
must yield the same result. This requires that
(2.2.33)
(2.2.34)
We have constructed modal E and H fields with real valued transverse
components and imaginary z-components. The general implications for a
24
propagating mode are that the tangential components of its E and H fields
are in phase, that their z-components are also in phase, and that the tan-
gential and z-components are 90° out of phase.
The fields of a forward evanescent mode vary as exp( -az). Here the
time reversal operation produces again a forward wave. Uniqueness requires
that these two waves be identical; and because of (2.2.29), this requires that
Ev = E~ Hv = -IT;; (2.2.35)
We have constructed an evanescent mode with a real valued electric field
and an imaginary magnetic field. In general, Ev and Hv of an evanescent
mode are 90° out of phase.
which yields
25
\7t • (Ev x ~ +~ X Hv)t - j({3v - (3f.l)
X (Etv X Htf.l + Etf.l X Htv)z =0 (2.2.41)
Here we have, again, separated the transverse (t) and the longitudinal (z)
components and used the transverse del operator \7t. The next step is to
integrate (2.2.41) over a cross-section z = const of the waveguide. Applying
the divergence theorem to the first term we get
II .f
+00
dx dy \7 t • 9 = ds 9 • et (2.2.42)
-00 c
where
(2.2.43)
and the line integral extends over an infinitely large curve enclosing the
waveguide, with et being a unit vector perpendicular to that curve. It is
easy to see that this line integral vanishes if at least one of the two modes is
a guided mode with fields decaying exponentially towards infinity. The line
integral also vanishes when both modes are radiation modes; the argument
to show this is somewhat more complicated and involves the oscillatory
nature of the radiation modes [2.3].
The terms remaining after integration are
II
+00
where we have dropped the factor ({3f.l - (3v) as we assume that {3vr={3w
The z-reversal symmetry allows a further simplification. In order to achieve
this, we apply (2.2.44) to a backward traveling mode (labeled -v) instead
of the corresponding forward mode (labeled v). According to (2.2.31,32),
the fields of the backward-traveling mode are given by
Et,-v(x, y) = Et,v(x, y) (2.2.45)
II
+00
II
+00
26
2.2.6 Mode Expansion and Normalization
Et(x,y) = LLavp,EvJ.!(x,y)
v p,
JJdvdp,a(v,p,)E(v,p,;x,y)
0000
+ (2.2.49)
o 0
Ht(x,y) = LLavp,HvJ.!(x,y)
v p,
JJdvdp,a(v,p,)H(v,p,;x,y)
0000
+ (2.2.50)
o 0
The summation here extends over the discrete and finite set of guided
modes and the integration extends over the continuous spectrum of ra-
diation modes. The discrete spectra in the above expressions are similar to
those encountered in hollow metal waveguides and the continuous spectrum
is similar to the angular spectrum of plane waves of a field in free space. It
is thus natural to call the continuous labels v and p, "spatial frequencies"
and to assign to them the dimension [cm -1 ]. In order to preserve the same
dimensionality for the field distributions of both the guided and of the radi-
ation modes, we have to keep the discrete coefficients avp, dimensionless and
assign the dimension [cm 2] to the coefficients a( v, p,) of the continuum. It is
advantageous to use only positive spatial frequencies; their lower limit, writ-
ten as 0 in the above equations, actually depends on the particular choice
of the labels v and p,.
In (2.2.49,50), we have not explicitly indicated the necessary summa-
tion over modes of different polarization (e.g., TE and TM modes) and over
degenerate modes with the same spatial frequency (e.g., the even and odd
radiation modes to be discussed in Sect. 2.3).
It is convenient to normalize the modal fields by means of the cross
power p(v,v,p,,"ji) such that
27
11
+00
11
+00
for continuous modes, where 8(v - v) is the delta function. These normal-
izations express, of course, also the orthogonality demanded by (2.2.48).
When the modes are not normalized, the cross-power P of the contin-
uous modes is given by
P(v, v, /1, Ii) = PvJ.t8(v - v)8(/1- II) (2.2.53)
11 11 dxdyE~J.t
+00 +00
11
+00
11
+00
=2 dxdyE*(v,/1) X Ht (2.2.55)
-00
11
+00
P = dxdy(Et X m + Et X Ht)z
-00
11
00
28
where P is measured in W. We identify aVJ.la~J.l as the power carried by a
discrete mode, and a( v, /-l )a* (v, /-l) as a spectral power density, which implies
that a power a( v, /-l )a* (v, /-l )L1v L1p is carried in the spatial frequency band
L1v L1/-l of th continuous mode spectrum.
In the above discussion, we have noted a formal analogy between the
discrete and the continuous modes. In the following, we shall take advantage
of this to simplify our notation by writing av for aVJ.l and a(v, /-l), Ev for
EvJ.l and E(v, /-l), etc., and by writing
(2.2.57)
(2.2.59)
where a v are the coefficients of the forward waves and bv are the coefficients
of the backward waves. With the help of the orthonormality relations we
determine the coefficients as
11
+00
= 11
+00
as expected.
Let us now consider evanescent modes which have imaginary propa-
gation constants. As none of these modes can carry power by itself, we
obtain cross power products. which are imaginary, and we have to change
the orthonormality relation to
11
+00
29
where the occurence of the + or - sign depends on the particular guide
configuration and its mode solutions. Keeping the mode expansion in the
form of (2.2.58,59), we obtain for the coefficients of the evanescent modes
11
+00
11
+00
which reflects the fact that an evanescent wave cannot carry power by itself,
but a combination of forward and backward evanescent waves can lead to
tunneling of power through a short region.
an similarly for 6H. If we insert this in the above theorem, we obtain the
relation
\7t· 9 - j6f3v' Sz = -j[6(w€)Ev' E:, + 6(wf-L)Hv' H';,] (2.2.70)
where \7 t is the transverse del operator,
S = Ev X H';, + E:, X Hv (2.2.71)
30
is the time averaged Poynting vector, and
9 = e:, X oHII + oEII X ~ - jo(311 • zS (2.2.72)
11 dxdy[o(w.::)EII·E~+o(WJL)HII·~]
+00
O(3I1· P = (2.2.73)
-00
where
(2.2.74)
-00
is the power carried by the mode. This is the variation theorem for dielectric
waveguides. One of the applications of this theorem is the determination of
the change L1(3 of the propagation constant due to a perturbation L1.::(x, y)
of the dielectric constant of the waveguide. Here we have Ow = 0 and 0JL = 0
and the theorem yields
11
+00
11
+00
P = dxdySz (2.2.77)
-00
The time-average density of the energy stored by the modal fields is given
by
d(w.::) * d(wJL)
w(x, y) = --a;;;-EII • Ell + -a;;;-HII • ~ (2.2.78)
31
which is written here in a form that allows for dispersive medium constants
t:( w) and f-L( w) [2.28]. The energy W stored per unit guide length is obtained
by integrating w over the guide cross section
11
+00
W= dxdyw(x,y) (2.2.79)
00
The group velocity Vg of the mode is the velocity at which signals carried
by the mode propagate. It is determined from the dispersion relation w((J)
of the guide by
Vg = dw/d(J (2.2.80)
Applying the variation theorem (2.2.73) to cases where all variations are
caused by perturbations in w alone, we deduce the simple relation
P = (dw/d(J)W = VgW (2.2.81)
11 11
+00 +00
wt = 11 dxdYf-LHt·II; 11
+00 +00
Wf = dxdYf-LHz·H; (2.2.83)
-00 -00
Here we have also distinguished between the energy portions stored by the
transverse (t) and longitudinal (z) field components, and have left out the
mode label 1/ to simplify the notation. Forming dot products of the modal
differential equations (2.2.16-18) with the appropriate field components and
combining the results one obtains
~t· E z X II; - j(Je z • Et X II; = jwt:Ez ' E; - jWf-LHt • II; ,(2.2.84)
When we integrate this over the guide cross-section, we find, as in Sect. 2.2.5,
that the first terms vanish; This leads to the simple relations
(JP = 2w(Wt - W;) (2.2.86)
32
Subtracting these two relations we find
W~ = W{ + W; = Wi + Wf = WJL (2.2.88)
stating the equality of the stored electric energy W~ and the stored mag-
netic energy WJL. The same relation also follows from the complex Poynting
theorem.
By adding (2.2.86 and 87) we obtain another interesting relation, namely
where (2.2.89)
This expression relates the phase velocity vp of the mode to the power flow P
and the quantity (Wt - W z ), which can be identified as the electromagnetic
momentum flow in the waveguide [2.27]. It should be contrasted to the
relation (2.2.81) for the group velocity Vg. Combining these two relations
we get
vp Wt + Wz (2.2.91)
Vg = Wt - Wz
(3 = X H - H*. V' X E]
-00
/ J dxdy(E
+00
X H* +E* X H) (2.2.96)
-00
J dx dy .:1EE* • Ej P
+00
.:1(3 =w
-00
J dx dy(V' J dx dy p,oEE*. E
+00 +00
35
Fig. 2.14. Sketch of an "asymmetric" slab
waveguide and the choice of the coordinate
system. Note that the z-axis lies in the film-
substrate interface
z
index nco This structure has also been called the "asymmetric" slab guide.
The modal fields can be derived from the wave equations of Sect. 2.2.3, and
the corresponding solutions have been discussed by McKenna [2.24]' Tien
[2.17], Marcuse [2.2], and others. We follow essentially Marcuse's treat-
ment, but cast the results in a simple form which employs the peak val-
ues of the fields in substrate, film and cover as well as the phase shifts at
the film-substrate and film-cover boundaries that play an important role
in the zig-zag wave picture. We have to distinguish between modes of TE
(Sect. 2.3.1) and modes of TM polarization (Sect. 2.3.2). Another distinction
is between guided modes and radiation modes, the latter divided into the
categories of substrate radiation modes, substrate-cover (also called "air")
radiation modes, and evanescent modes. Section 2.3.3 deals with multilayer
waveguides.
In accordance with the wave equation we define various transverse de-
cay (,i) and propagation constants (Ki) by
(2.3.1)
2
Kf = nf2k 2 - f32 (2.3.2)
(2.3.3)
where the subscripts s, f, and c refer to substrate, film, and cover respec-
tively.
Table 2.3 shows the ranges of propagation constants f3 corresponding to
the various mode types and categories, and also the associated ranges of the
transverse propagation constant Ks in the substrate. Because of its conve-
nient range, the latter is chosen as the spatial frequency for the continuous
mode spectrum.
As we are dealing with a planar problem, all field solutions can be made
independent of the y-coordinate.
36
Table 2.3
Modes (3 lis
2.3.1 TE Modes
Hx = -(f3/w/-l)Ey (2.3.5)
Hz = (j/w/-l)fJEy/fJx (2.3.6)
with the Ey component obeying the wave equation
fJ2Ey/fJx 2 = (f32 _n 2k 2)Ey . (2.3.7)
The boundary conditions of (2.2.8) demand that Ey (and thereby auto-
matically Hx) and fJEy/fJx (and thereby Hz) be continuous across the film
boundaries at x = 0 and x = h.
For guided modes we have
Ey = Ecexp[ -1'c(x - h)] for h<x (cover)
Ey = EfCOS(KfX - <Ps) for O<x<h (film)
Ey = Esexp( 1'sX ) for x<O (substrate) (2.3.8)
Application of the boundary conditions yields the formulas for the phase
shifts
tan <Ps = 1's/ Kf , tan <Pc = 1'ci Kf , (2.3.9)
and the dispersion relation
Kfh ~ <Ps - <Pc = V7r (2.3.10)
where the mode label v is an integer. This is in agreement with the dispersion
relation obtained from the zig-zag wave picture (Sect. 2.1). We also get a
relation between the peak fields E s , Ef and Ec of the form
f f _ N ) = E2(n
E2(n 2 2
sf _
2 ns2) = E2(n 2 2
c f _ nc ) (2.3.11)
where N = f3 / k is the effective refractive index.
In the above form, the modes are not normalized for power. We calcu-
late the power P carried by a mode per unit guide width as follows
37
J J
+00 2{3 +00
f!
-00 -00
0 2
= N -Ef • heff = EfHf· heff , (2.3.12)
J-Lo
where
1 1
heff=h+-+- (2.3.13)
'Ys 'Yc
is the effective thickness of the waveguide, as discussed in Sect. 2.1.4.
For the substrate radiation modes, the field distribution is
Ey = Ecexp[ - 'Yc(x - h)] for h< x
Ey = Er cos [lI:rC x - h) + 4>cl for 0<x <h
Ey = Es cos(lI:sX + 4» for x <0 . (2.3.14)
2 2 f n2 - ns2 2 ]
Es = E f [ 1 + n~ _ N2 sin (4)c -lI:fh) (2.3.18)
Here we have no dispersion relation leading to discrete values for (3, and we
choose II:s as the independent continuous variable.
The above field of a substrate mode is exactly the same as that created
by a plane wave incident from the substrate side, with II:s = kns cos Os used
as a measure for the angle of incidence Os. The incident wave is refracted and
partially reflected at the film-substrate boundary and totally reflected at the
film-cover boundary. The phase shift incurred at that total reflection is 24>c,
and the phase-shift for reflection from the film-cover combination is 24>.
Interference between the incident and reflected waves creates the sinusoidal
standing-wave patterns in film and substrate.
For the cross power P(lI:s, its), needed to normalize the substrate modes,
we calculate
J
+00
P = -2 dx Ey(lI:s)H:x;(its)
-00
(2.3.19)
38
The substrate-cover radiation modes are degenerate; we obtain two
independent field solutions for each given K,s. Great care must be exercised
to select two solutions which are orthogonal to each other, as required by
the mode-expansion formalism. A convenient choice are modal fields which
become even or odd functions of (x - h/2) in the limit of a "symmetric"
waveguide where ns = nco For simplicity, we also call these modes "even"
and "odd" in the asymmetric case. Their fields are
(2.3.22)
and for the even modes
K,s tan 4>s = K,f tan 4> , (2.3.23)
(2.3.28)
For the cross power between even or between odd modes we calculate
(2.3.29)
39
When needed, the second delta function in this expression can be rewritten
as
8( Kc - ~c) = (KC/ Ks)8( Ks - ~s) (2.3.30)
which becomes a condition for the phase shift <p of the form
2.3.2 TM Modes
Section 2.2.3 gives the relations for the TM modes
Ey = Hx = Hz = 0 (2.3.36)
Ex = (f3/wc)Hy (2.3.37)
Ez = -(j/wc)oHy/ox (2.3.38)
where v is an integer. We also get a relation between the peak fields of the
form
(2.3.45)
For the power per unit guide width carried by a mode we calculate
P =2 J dx
+00
-00
Ex Hy = WED
-
2(3
J dx Hy / n
+00
-00
2 2
(2.3.46)
(2.3.52)
41
Comparing the above expression with the corresponding formulas for the
TE modes, we note the appearance of the reduction factors qs and qe.
As in the TE-mode case, we use "'s as the continuous-mode label. For
the cross power P( "'s, Ks) we calculate
J dx Ex ("'s)Hy(Ks)
+00
P =2
-00
(2.3.53)
(2.3.54)
where </> is the same for the even and odd modes. The relations between the
phase shifts, as derived from the boundary condit"ions, are
("'s/n;)cot </>s = ("'r/nl)cot </> , (2.3.55)
(2.3.60)
(2.3.61)
(2.3.62)
The cross power between even or between odd modes of different "'s is
42
(2.3.63)
This translates into a condition for ¢ which has the same form as (2.3.34)
for the TE mode but with Kf replaced by Kr! nl,
Ks by Ksl n;,
and Kc by
Kc/n~.
1
nc
nSf \A. a.
waveguide with substrate index ns and cover
index nc. The z-axis indicates the direction
of mode propagation
43
analysis 6f this guide will concentrate on the TE modes. At the end of this
section we will give a simple substitution rule that leads to results for the
TM modes.
The theory of multilayer stacks [2.47,48] starts with (2.19-21) and defines
two field variables U and V by
U=E y V =wpHz (2.3.65)
which describe the transverse variation of the optical field. These definitions
are chosen, because U(x) and Vex) are quantities that are continuous at the
layer boundaries. From (2.19-21) we obtain the relations
U' = -jV (2.3.66)
(2.3.67)
where
11: 2 = n2k2 - (32 (2.3.71)
as before, see e.g. (2.3.1-3). The constants A and B can be replaced by the
input values Uo = U(O) and Vo yeO) at the input plane x = 0 of the
layer. We have
(2.3.74)
44
",here the pairs (Uo, Vo) and (U, V) have been written as vectors, and M is
the characteristic matrix of the layer. It has the form
M = IjKcos( K(X))
sin KX
(j/ K) sein(K) x)
cos KX
I (2.3.75)
Mi = IjKiCOS(K(ihhi))
sin K i
(j/Ki)s(in(hK,).h i )
cos Ki i
I (2.3.76)
where
K?-n?k
• -,
2 _f.l
jJ
2 (2.3.77)
(U~-I)
Vi-I
= Mi (U~)
Vi
. (2.3.78)
(2.3.79)
M == Imll
m21
mI21 = Ml . M2 . M3 ... Mn
m22
(2.3.80)
where mll, mI2 etc. are the matrix elements which we will use below.
On the cover side, we have light with a transmitted amplitude A e , and the
corresponding output variables
(2.3.83)
where
K~ = n~ k2 - (32 = -,.;; (2.3.84)
which obey the transverse wave equation (2.3.68). Instead of the transverse
propagation constants Ks and Ke , we must now use transverse decay con-
stants "Is and "Ie, which are already defined in (2.3.81 and 84). We postulate
that the fields should decay away from the multilayer, and obtain for the
input and output field variables
Uo =As (2.3.89)
46
Dividing these two relations leads to the desired dispersion relation for a
multilayer slab guide
(2.3.91)
It is expressed in terms of the decay constants IS and IC, and the elements
of the characteristic matrix of the stack. The relation establishes the connec-
tion between the frequency w = kc of the light and the propagation constant
(3 of the mode guided by the multilayer. It is valid for an arbitrary number
of layers and can also be used for the approximate analysis of waveguides
with graded index profiles. As a useful illustration we shall discuss below
the examples of the four-layer and the symmetric five-layer waveguides.
Four-Layer Waveguides
Our first example is a four-layer slab guide such as that used in an LOC
laser structure. A sketch of this guide is shown in Fig. 2.16 together with
the key guide parameters. For simplicity, we choose a substrate and cover
of equal index (ns = n c ), which implies equal decay constants IS = IC = I.
For this case, the dispersion relation simplifies to
(2.3.92)
This relation implies the modal cut-off condition m21 = 0 which corresponds
to the case of zero decay (, = 0) of the fields in the substrate. The required
elements of the characteristic matrix M = M1M2 of the two-layer stack
are obtained from (2.3.76) by matrix multiplication with the result
ml1 = cos( I'qh1) cOS(1I:2h2) - (11:2/11:1) sin(1I:1h1) sin(1I:2h2)
m22 = cos(1I:1h1) cOS(1I:2h2) - (11:11 11:2) sin(1I:1h1) sin(1I:2h2) (2.3.93)
47
[COS(Klhl) - (KIll) sin(Klhl)][ COS(K2h2) + (-r/K2) sin(K2h2)]
= -[ COS(KIh l ) + (iIKl) sin(Klhl)]
X [COS(K2h2) - (K211) sin(K2h2)] (2.3.94)
This agrees with the dispersion relation obtained by other methods [2.6,41].
As a further check, we set h2 = 0 and find the result to agree with the
dispersion relation of the symmetric 3-layer slab guide as expected.
z
Fig.2.17. Symmetric 5-layer guide. The
heights of the sandwiched films are hI,
2h2, and hI, and their indices are nI, n2,
and nI as shown
where we use the lower case to distinguish the matrix elements of the pairs
48
(mll,mI2 etc;) from the matrix elements of the 3-layer stack. Expressions
for the elements mll, m12 etc. are given by (2.3.93). Inspecting these ex c
pressions, we find that the matrices of the two-layer pairs are related by a
reversal of the elements mll and m22, as reflected in the equations above.
Performing the matrix multiplication, we find the following relations
Mll = M22 = mll m22 + m12 m 21
M12 = 2mllm12 M21 = 2m21m22 (2.3.98)
This contains the relation Mll = M22 which we expect for a symmetric
stack.
With the latter simplification, the dispersion relation of the 5-layer
guide becomes
2j,Mll = M21 - ..l M12 (2.3.99)
(2.3.102)
Because det (MIM2) = 1, the two relations cannot be simultaneously sat-
isfied for the same wand (3 values. They correspond to different sets of
modes. Closer inspection shows that these sets are the even-order and odd-
order modes.
Equation (2.3.101) corresponds to the even-order modes. If we insert
the matrix elements from (2.3.93) and sort the terms with common functions
of (K2h2), we obtain
K2 sine K2h2) . [cos( Kl hI) + (, / Kd sin(Klhl)]
= Kl cos( K2h2)[(')' / Kl) cos( Klhl) - sine Klhl)] (2.3.103)
Note that sorting for common functions of (Klhd yields the dispersion rela-
tion given in [2.45]. However, we proceed with (2.3.103), apply well known
identities for trigonometric functions and arrive at the result
(2.3.104)
This agrees with the dispersion relation for even-order modes obtained by
a different method [2.6].
For the odd-order modes we start with (2.3.102), proceed along similar
lines as before, and obtain the dispersion relation
(2.3.105)
49
TM Modes
The preceding discussion of multilayer guides is valid for the TE modes. We
can modify the TE results to obtain results for the TM modes with the aid
of a simple substitution.
For TM modes we have Ey = Hz = Hx = O. The analysis starts with
(2.2.24-26) and follows a parallel path to the TE analysis. One defines the
field variables
U=H y v = wcoEz (2.3.106)
This already gives us the key for the required substitution. A comparison
with equations (2.3.69,70) shows that the general TM solution is the same
as the general TE solution if we substitute
(2.3.109)
and similarly for f'
Note, that this substitution should not be applied to the phase terms
(lI:x), (lI:lhI) etc., as these are identical in the two general solutions.
The substitution can be applied to all results that follow from the gen-
eral solution, i.e. almost all results of interest. This includes the formulas for
the reflection and transmission coefficients, and, particularly, the dispersion
relation. For TM modes, the latter assumes the form
-j(mllfS/n; + m22fcln~) = m21 - ({sfcln;n~)m12 (2.3.110)
as a result of the substitution. If we apply the substitution to the char-
acteristic matrix of a single layer, as given in (2.3.76), we obtain the TM
result
M· _I
1 -
cOS(lI:ihd
.,...j(lI:dn~) sin(lI:ihd
-j(n~/lI:i)sin(lI:ihd 1
cos( lI:ihi) .
(2.3.111)
In similar fashion, we can adapt the other TE results obtained above to the
case of TM waves.
50
2.4 Planar Guides with Graded-Index Profiles
51
The index profile of (2.4.2) corresponds to the potential well of the harmonic
oscillator (see, e.g., [2.29]), and the solutions of the wave equation are
Ey = H v (v'2x/w)exp(-x 2 /w 2 ) , (2.4.4)
where the Hv are the Hermite polynomials defined by
dV
Hv(x) = (-ltexp(x2) dx vexp( _x 2 ) (2.4.5)
For the lowest orders we have
Ho(x) = 1
Hl(X) = 2x ,
H2(X)=4x 2 -2 ,
H3(X) = 8x 3 -12x (2.4.6)
The Hermite-Gaussian functions of (2.4.4) are also known as the standard
description for the modes of laser beams and laser resonators [2.3,49], where
the parameter w is called the "beam radius". The latter is given by
(2.4.7)
and indicates the degree of confinement of the fundamental mode. For the
propagation constant f3v and the effective index N v of a mode of order v,
one obtains
f3~ = n;k2 - (2v + l)nfk/xo (2.4.8)
(2.4.9)
52
Fig.2 .19. The "l/coshz " index profile. The
nominal guide thickness h is indicated
I
I
I
I
I
I
I
I
I
I
I
-i------~1----+---~I~--------~X
I-h~
I
The solution of the wave equation (2.4.1) for the "1/cosh 2" profile of (2.4.10)
yields a very simple expression for the maximum mode number 8 ~ v of
guided (bound) modes supported by the waveguide, which is
(2.4.13)
For the propagation constants f3v and the effective indices Nil we obtain
f3~ = n;k2 + 4(8 - v? /h2 , (2.4.14)
N; = n; + (8 - v)2(>'Irrh)2 (2.4.15)
where the U II are hypergeometric functions [2.28]. For even mode number v
we have
UII =1- !V(28 - v)sinh 2(2x/h)/(1 01!)
+ tv(v - 2)(28 - 7/)(28 - V - 2)sinh4(2x/h)/(lo 3 02!) + '" ,
(2.4.17)
and for odd v we get
53
For the lower mode orders these functions simplify to
Uo = 1
Ul = sinh(2x/h)
U2 =1- 2(s -1)sinh2(2x/h)
U3 = sinh(2x/h)[1- ~(s - 2)sinh2(2x/h)] (2.4.19)
This is also a special case of the Poschl-Teller potential for which exact
solutions are available, as discussed in detail by Gordon [2.51].
The exponential profile is another case for which exact solutions are available
[2.52]. We adapt these here to the symmetric profile sketched in Fig. 2.20
and described by
n 2 (x) = n; + 2n s.dnexp( -2Ixl/h) (2.4.21)
54
The solutions for the modal fields of this profile can be expressed in terms
of Bessel functions Jp of the first kind and of non-integral order p,
Ey = Jp(Vexp[ - x/h]) for x >0
= Jp(Vexp[x/hD for x <0 (2.4.24)
The order p// is determined by matching the solutions at the boundary x =0
for given V. For the even modes, we require that
J~(V) =0 (2.4.25)
and for the odd modes we have to demand
Jp(V) =0 (2.4.26)
Each of the two conditions yields approximately V /7r discrete solutions for
p//. In terms of p//, we can write for the propagation constant /3//
/3~ = n;k2 + p~/h2 , (2.4.27)
0.5
c:
<ItI)
c:
C\J 0.4
..........
N
tI)
c: 0.3
I
N
Z
N " 0.2
>
"-
c...
0.1
v =kh.j2n 56n
Fig. 2.21. Normalized w-f3 diagram for planar guides with an exponential profile. Shown
are the dispersion curves for the modes of odd order which correspond to the modes of
guides with an asymmetric exponential profile of mode number // = //asym, as given by
(2.4.30). (After [2.53])
55
(2.4.29)
for the odd-order modes, which is taken from Carruthers et al. [2.53] and
represents the results of a numerical solution of (2.4.26).
x I
x
, I
, I
, I
\ I
" ...
\ /
_-"" /
Fig. 2.22. Index profile with strong asymmetry (solid curve) and the associated sym-
metric profile (dashed curve). The corresponding approximate field distribution E( x) is
also shown
56
using the simple relation
2vasym +1 = Vsym (2.4.30)
between the mode numbers vasym of the asymmetric profile and the mode
numbers Vsym of the corresponding symmetric profile.
Even though the actual field at the film-air interface is very small, it is
not exactly zero. For some applications, such as the design of guided-wave
filters with a surface corrugation, one needs to know the field values at the
interface. Haus and Schmidt [2.55] have described how one can improve
the above approximation and obtain estimates for these field values. They
assumed evanescent fields in the cover region with the decay constant Ie
given as usual by
,~ = f32 - n~k2 (2.4.31)
Matching the fields at the boundary x = 0 they obtain for the y-component
of the electric field
Ey(O) = -1 (dEy)
- (2.4.32)
Ie dx x=o
which relates the field at the surface x = 0 to the slope dE / dx . The latter
is calculated from the approximate mode solutions discussed above.
Exact solutions for the asymmetric exponential profile were discussed
by Conwell [2 .52].
57
mathematically by
n(xd = N = {3/k i = 1,2 (2.4.33)
The values of {3 which obey the condition
J
X2
with integer v, are the propagation constants (3v predicted by the WKB
method. In terms of the effective index, this condition can be written as
A
J
X2
dxVn 2 - N2 = 2(v +!) (2.4.35)
For the modal fields, the WKB method predicts oscillatory field distribu-
tions in the region where n(x) > N (i.e., Xl < X < X2) and exponentially de-
caying fields where n(x) < N (i.e., x> X2 and X < xI) .
For the special case of the parabolic profile, the WKB predictions are
known to agree exactly with the closed form solutions (2.4.8) available for
this profile.
The planar optical guides discussed in the preceding two sections provide
no confinement of the light within the film plane, i.e. the y-z plane. There,
confinement takes place in the x-dimension only. Optical channel guides can
provide this additional confinement in the y dimension. Channel guides are
used in many active and passive devices of integrated optics, including lasers,
modulators, switches and directional couplers. The additional confinement
can help to bring about desirable device characteristics such as savings in
drive power and drive voltage. In addition, it is required for the design of
single-mode structures that are compatible with single-mode fiber guides.
The following subsections will give an overview of channel guide ge-
ometries, discuss the vector wave equation suitable for the exact analysis
of channel gcides, and discuss methods of approximation including the sep-
aration of variables, the method of field shadows, and the effective index
method.
Figure 2.24 shows sketches of the x-y cross sections of six different types of
channel guides. For simplicity, the figure shows abrupt transitions of the re-
58
Fig.2.24a-f. Cross sections of six
channel guide structures
59
for this analysis are the Maxwell's equations (2.2.6,7) for the complex am-
plitudes, which we repeat here, for convenience,
\7 X E= -jw{loH , \7 X H=jwe:E . (2.5.1)
In the following we present a derivation which will yield the vector wave
equation for the transverse fields of a channel guide [2.4,6, 7]. As a first
step, we take the curl of Maxwell's equations, which gives
\7 X \7 X E = -jW{lO \7 X H = w2 e:{loE , (2.5.2)
The second of these divergence relations can be written in the more conve-
nient form
\7. E = -E· \7 In e: (2.5.7)
We note that the transverse index gradients described by \71ne: adds terms
to the usual wave equation which tend to couple the components of the
vector fields E and H. Closer inspection shows that the longitudinal field
components are decoupled from the transverse components. This becomes
clear when we separate the fields into transverse (Et, Ht) and longitudinal
(Ez, Hz) components as in Sect. 2.2, and write the modal fields in the form
E = (Et + Ez)exp( -j{3z) H = (Ht + Hz)exp( -j{3z) , (2.5.11)
omitting the modal subscript v for reasons of simplicity. With this we obtain
the vector wave equations for the transverse fields
60
\l2 Et + \l(Et· \lIne:) + (w 2e:f-lO - f32)Et =° (2.5.12)
We note here that the quantities e:, Et and Ht are independent of z. There-
fore, we have \It lne: == \lIne:, \ltEt == \lEt, and \It X H t == \l X H t , etc.,
where \lt is the transverse grad operator used earlier. We notice that the
gradient term in the vector wave equations will generally couple the x- and
y-components of the fields. However, no longitudinal field components ap-
pear in these equations.
In principle, we need to solve only one of the two vector wave equations
for Et or Ht. The corresponding z-components follow from the divergence
relations (2.5.6 or 7), which can be expressed as
jf3Ez = \l. Et + E t • \lIne: (2.5.14)
jf3Hz = \l. H t (2.5.15)
A variety of methods have been reported which are suitable for the numer-
ical analysis of channel waveguides based directly Maxwell's equation for
the guide [2.6,5]. For the case of buried rectangular guides, Goell [2.58] has
employed cylindrical space harmonics to analyze guides with aspect (width
to height) ratios between 1 and 2. Schlosser and Unger [2.59] have de-
scribed a numerical method which is suited to rectangular guides with large
aspect ratios. Additional numerical results for rectangular guides have been
reported in [2.60,61], and the use of variational techniques for the analysis
of these guides has been presented in [2.34,39]. For a recent detailed review
of numerical methods the reader is referred to [2.81].
Figure 2.25 shows dispersion curves obtained by Goell for buried guides
of index nf surrounded by a cladding of index ns. For small index differences
(nr - n s ), the results can be plotted in broadly applicable normalized form
as shown.
Here, we have used the same normalizations as those used for planar
slab waveguides in Sect. 2.1, where
(2.5.16)
is the normalized guide index related to the effective index N, and
(2.5.17)
is the normalized guide thickness (or height). The curves are labeled with
the aspect ratio, i.e., the ratio between the guide width wand the guide
61
1.0
fJrnf
n.
~
h
N.
c
0.8
...
I
N
c
.... 0.6
.....
N.
c 0.4
I
N
z 0.2
II
..0
2 4 6 8 10 12
V=kh(nl-nl>1/2
Fig.2.25. Dispersion curves of a buried channel guide of height h and width w. The
normalized guide index b is shown as a function of the normalized frequency (or guide
thickness) V for the w/h ratios of 1,2, and 00, (After [2.58])
62
Here it is required that X and nx are functions of x only, and Y and ny are
functions of y only.
Under these conditions, the two-dimensional wave equations can be
separated into two one-dimensional parts
J2x + (k 2 nx2 - /3x)X
-2
2
=0 (2.5.21)
dx
(2.5.22)
These equations are now in a form which allow us to make use of known
planar slab guide solutions, such as those given in Sect. 2.4. Once the solu-
tions X, /3x, Y and /3y are determined, the propagation constant /3 of the
channel guide is given by
/3 2 = k2n5 + /3; + /3; (2.5.23)
n2 ( x, y) = n; (1 -
x2 y2)
2" - "2 (2.5.24)
Xo Yo
where Xo and Yo indicate the guide height and width, respectively. The
modal properties of this guide follow directly from the properties of the
parabolic slab guide discussed in Sect. 2.4.1. The fundamental mode is a
Gaussian 2 2
Et(x,y) = Eoex p ( - - :~ ~~) (2.5.25)
(2.5.26)
As in (2.4.4), these parameters are also used for the Hermite-Gaussian func-
tions which describe higher-order modes. The propagation constants of the
modes of the guide are obtained from
/3 2 = k 2 n; - (2v + l)nfk/xo - (2p, + l)nfk/yo (2.5.27)
- ---+----.... y
•t
h
Fig. 2.27. Method offield shadows. The sketch shows the x-y cross section of a composite
guide made up by summing the permittivities (n 2 ) of an x-slab guide of height h and a
y-slab guide of height w. The various n 2 values are indicated
of the fields X(x) of the x-slab guide and Y(y) of y-slab guide. The prop-
agation constants j3x , j3y and effective indices N x , Ny of the slab guides
determine the corresponding quantities 13 and N of the channel guide via
N 2 = N x2 +Ny2 (2.5.30)
(2.5.31)
(2.5.32)
where V is · the normalized height of the channel guide. Fig. 2.8 is used to
determine the normalized guide indices bx(Vx ) and by(Vy) ofthe slab guides.
65
Now we invoke the definition of the normalized indices
b - N x _ n s + n f j2
2 2 2
(2.5.33)
x- nl- n~
(2.5.34)
(2.5.35)
to derive the surprisingly simple relation for the normalized index b of the
channel guide
b = bx + by -1 (2.5.36)
This means that the normalized dispersion chart for slab guides shown in
Fig. 2.8 can be used in a simple manner to determine the normalized guide
index of a buried channel guide.
For the special case of a square channel where bx = by we find
b = 2b x -1 (2.5.37)
We can use this relation to relabel the vertical axis of the slab-guide disper-
sion chart to obtain the dispersion chart for the square channel guide.
The variation theorem (2.2.75) can be used to obtain a correction L1f3 for
the propagation constant due to this error. We get
66
2.5.6 The Vector Perturbation Theorem
Channel-guide results obtained via the scalar-wave approximation can be
improved with the vector perturbation theorem. This is quite useful in cases
where small differences between two parameters play an important role. An
example is the determination of the birefringence of a channel guide.
In order to derive this theorem, we assume that the scalar wave results
Eo(x,y) and 130 are known. They obey the scalar wave equation (2.5.18)
V 2 Eo + (n 2k 2 - f35)Eo = 0 (2.5.39)
The exact solutons E( x, y) and 13 must obey the vector wave equation
(2.5.12)
(2.5.40)
If Eo is a good approximation, we can write
E= Eo +El (2.5.41 )
and treat the quantities El, Vln e, and L1f32 as small perturbations, where
(2.5.42)
Subtracting (2.5.39) from (2.5.40) and dropping perturbations of second
order yields
V 2 El + V( Eo . Vln e) + L1f32 Eo + (n 2k 2 - f32)E 1 = 0 (2.5.43)
We form the dot product of this with Eo and the dot product of (2.5.39)
with El and subtract the results. We proceed by integrating over the guide
cross section and employing Green's theorem to obtain
L1f32 = Jdx dy Eo· V(Eo . Vlne)/ Jdx dy E5 (2.5.44)
L1fJi-E = 2
1 Jdx dy '1jJ2 8 2
8y2ln c / Jdx dy '1jJ2
(2.5.48)
This leads to the following formula for the geometrical birefringence of the
channel guide
2 - fJTM
fJTE 2 = 21 Jdx dy'1jJ 2(
2
8 2 - 8y2
8x
2
8 ) In c / Jdx dy '1jJ 2 (2.5.49)
Inserting this into the theorem (2.5.48) we obtain directly the result
2 1 2 1
L1fJTE = -2" L1fJTM = - 2 (2.5.51)
Yo Xo
which agrees with the ret>ults obtained by other methods [2.4,40,75]. For
the birefringence of the parabolic channel we get the estimate
68
2.5.7 The Effective-Index Method
Because of its immediate intuitive appeal, the effective index approach has
been used since the beginnings of integrated optics. It has helped in the
understanding' of structures such as guided-wave prisms, lenses and gratings.
It has been proposed for the approximate analysis of channel guides by
Knox and Toulios [2.64], and has produced results which were in close
agreement with more exact computer results as well as experimental results
for a considerable number of practical guide structures. This includes its
application to ridge guides [2.65], buried channel guides [2.66] and diffused
channel guides [2.66].
The effective-index approach starts with a birds-eye view of a planar
film guide, with the film in the y-z plane. The guide is viewed in the x-
direction. For a uniform planar guide, the viewer sees a uniform effective
index N independent of y and z. When small variations are introduced ei-
ther in the guide thickness or the refractive indices, the viewer will see an
effective index pattern N(y, z). When this pattern has the shape of a famil-
iar bulk optical component, analogy arguments are used to understand its
characteristics. For a channel guide the viewer sees the pattern reminiscent
of a planar film guide with the film in the x-z plane. This is used to predict
the modal fields and the propagation constant~ of the channel guide.
In the following, we discuss the application of the effective-index method
to step-index channel guides. For a detailed treatment of diffused channel
guides, the reader is referred to Hocker and Burns [2.66]. Our discussion
aims to provide broad applicability while it uses the rib-guide structure as
an illustrative example (and gives, in parenthesis, numerical results for the
specific case of >.. = 0.8/Lm, nf = 2.234, ns = 2.214, nc = 1, h = 1.8/Lm,
1 = l/Lm, W = 2/Lm, a rib guide made of Ti: LiNb03). The normalized
guide parameters N, V and b, introduced in Sect. 2.1, will be used through-
out to allow for easy scaling of the results. As shown in Fig. 2.28, we use
the subscripts f and 1 to distinguish between the parameters of the channel
and the lateral guides, respectively. The figure shows the cross section (x, y)
of the rib guide example together with the top view (z, y) of the channel
guide. From this view we see a channel of width wand effective index N f ,
Vr = khJnl- n~ (2.5.53)
VI = kZJnl- n~ (2.5.54)
70
Table 2.4. Effective index parameters for channel guides
f) Vr = khJn; - n; N2r -- 2
nsl + br( nr2 _ 2)
n s1
::t Ridge (1 - b,)(n~l - n~2) + br(n; - n~l) beq (1 + br· aridge) + b,(1- beq )
V, = klJn; - n; Nl = n~2 + b,(n~l - TI:~2)
method. Here h always refers to the height of the channel, and I to the
height of the lateral guide. The following comments refer to specifics in the
treatment of the six structures (Table 2.4).
is used for most structures to simplify the expressions and to determine the
b parameters from the normalized plots of Fig. 2.8.
A similar simplification results from the introduction of a channel mea-
sure ach defined as
(2.5.60)
(2.5.62)
72
(2.5.63)
1.0
0.8
w/h=1
I
l----
E11~
~
~
-
0.6 V E12
~
l--
C\I
It
0.4
)~ ~
V
C
t,l'1
, 1 ~
~/
[7 ,
C\I 0.2
Z I ~
• o :~ I (a)
..c o 0.4 0.8 1.2 1.6 2.0 2.4 2.8 3.2 3.6 4.0
VI.". = Vf / .". = (2h/A) (nl- n,2)112
C\I
c
.. 1.0
I
w/h=2 L--- l - i.---
-
0.8
C\I
c
E11,
V
~
~
V Ii'"
~
J,...-- I -
17 ? ~E1&
~
V
c 0.4 Vol
j I ) . 'I
/II
0.2
!J'/
C\I
14' V
z
L.' I (b)
..c o
o 0.4 0.8 1.2 1.6 2.0 2.4 2.8 3.2 3.6 4.0
VI.". = Vf/.".= (2hlA) (nl- n,2)1/2
Fig.2.29a,b. Normalized dispersion curves for a buried channel guide comparing the
predictions ofthe numerical calculations (dot-dashed lines), of the effective index method
(solid lines), and of the field-shadow method (dashed lines). Comparisons are shown for
the aspect ratios of w/h 1 and w/h =
2. (After [2.66]) =
1 We are grateful to these author's for pointing out a misprint in their figures which
is corrected here.
73
modified for our purposes. The figure shows the normalized dispersion curves
b(V = Vr) for aspect ratios of w/h = 1 and w/h = 2 for the case of small
index differences. The solid curves show the prediction
b=bf·beq (2.5.65)
of the effective index method. The dot-dashed curves show Goell's data. As
a further comparison the figure shows also the prediction (2.5.36)
b = bz + by -1 (2.5.66)
of Marcatili's shadow field method depicted by the dashed curves. Data are
presented for the fundamental Ell mode as well as for the modes E12 and
E21 of the next higher order. The effective index method is seen to give
fairly good predictions even near cut-off. For larger aspect ratios we would
expect even better predictions because effects at the channel edges are then
of lesser importance.
74
guide deformations such as surface irregularities. This is followed by a listing
of the standard coupled-wave solutions both for co-directional and contra-
directional interactions, and finally we shall treat periodic waveguides and
present results for the specific example of a planar waveguide with periodic
surface corrugations.
\7 X H=jwE:E+jwP (2.6.2)
The following steps are now quite analogous to those used in the derivation
of the orthogonality relation in Sect. 2.2.5. We consider two different induced
polarizations PI and P 2 and the fields caused by them, and obtain
(2.6.3)
which is essentially the Lorentz reciprocity relation. Now we set P2 = 0
and identify the field 2 with a mode of the waveguide. As in Sect. 2.2.5, we
integrate over the guide cross-section, use the divergence theorem, and find
11
+=
-=
8
dx dy 8z (EI X H2 + E"2 X HI)z = -jw 11
+=
dx dy Pl· E2 . (2.6.4)
-=
The next step is to expand the transverse components of the field 1 in terms
of the modes of the guide according to (2.2.58,59)
EIt = :L) av + bv)Etv Hit = "L( av - bv)Htv (2.6.5)
where we use the L: notation of Sect. 2.2.6. It should be noted that the
coefficients av ( z) and bv ( z) have to be regarded as functions of z in the
present context. If we choose for the field 2 a forward running mode
(2.6.6)
insert these fields and the r.:!0de expansion into (2.6.4) and apply the or-
thogonality relation (2.2.51), we then find that the bv coefficients drop out
and we obtain
a~ + jf3ftaft = -jw 11
+=
dx dy p. EZ (2.6.7)
-=
where the prime indicates differentiation with respect to z. Similarly we get
75
11
+(X)
The above formulas hold only for propagating modes (real (3); for the evanes-
cent modes we have to use the orthogonality relation (2.2.63). It is usually
convenient to define the amplitudes AJL(z) and BI-'(z) of the forward and
backward propagating modes by
(2.6.10)
The change of these amplitudes due to the presence of sources can then be
expressed as
11
+(X)
11
+(X)
76
(2.6.15)
..::1c:. c:
= c:+ ..::1c: .L:(av - bv)Ezv (2.6.18)
where we have used the mode expansion (2.6.5) and the modal Maxwell
equation (2.2.16). We are now ready to insert the components of Pinto
(2.6.11 and 12) with the result
11
+00
11
+00
77
where we have used the symmetry relations (2.2.31) to express the mode
distribution E-J.L(x,y) in terms of the components of the field EJ.L(x,y) of
the corresponding forward running mode. To simplify these expressions,
we introduce tangential and longitudinal coupling coefficients KtJl(z) and
K~Jl(z) defined by
1f dx dy Llc:Etv' EtJl
+00
K~Jl = w (2.6.21)
-00
(2.6.22)
leading to real and positive quantities for positive .dc:. Using these coupling
coefficients and the mode amplitudes of (2.6.10) we can rearrange (2.6.19
and 20) in the final form
A~ = - j L: {Av(K~Jl + K;Jl)exp[ - j(,8v - ,8Jl)z]
+ Bv(K~J.L - K;Jl)expu(,8v + ,8Jl)z]} (2.6.23)
These two expressions form the basis for the solution of a number of coupled-
mode problems. They show the change in the amplitude of each mode (f-l)
as a function of the deformation .dc:, the modal field distribution, and of the
amplitudes of all other modes present in the guide. Depending on the partic-
ular problem at hand, one can usually make some simplifying assumptions
at this stage. A very common and usually good assmuption is that only
two guided modes are important and that all other modes can be neglected.
This leads to coupled-wave interactions with characteristics that are dis-
cussed further in the next subsection. Another common assumption is that
.dc:( x, y, z) is only a small perturbation of the dielectric constant c:( x, y) of
the waveguide. Often, this is also a good assumption, but there are config-
urations of interest where it is not justified. An example is the corrugated
glass waveguide used for filter devices where a corrugation of the glass-air
interface leads to a large .dc:. This is illustrated further in the discussion of
Sect. 2.6.4.
The expressions above point out a general difference between co-direc-
tional coupling (e.g., coupling between two forward modes AJ.L and Av)
and contra-directional coupling (e.g., between AJl and BJl) which occurs
in the presence of Ez components. As Kt and K Z have equal sign, the
factor (Kt + KZ) indicates stronger coupling for co-directional interactions
as compared to contra-directional interactions where we have the factor
78
(Kt - KZ). This happens because forward and backward modes have Ez
components of opposite sign and Et components of the same sign.
where I\: is the coupling coefficient which, in the simple case considered here,
is real and uniform (i.e., independent of z), and 8 is a normalized frequency
which measures the deviation from synchronism (for which one has 8 = 0).
In the next subsection, these two parameters are derived explicitly for the
example of a corrugated waveguide. By means of the simple substitution
A = Rexp(-j8z) B = Sexp(j8z) (2.6.27)
80
80 Fig, 2.30. Wavelength response of a
FILM CORRUGATIO"GRATING
corrugated waveguide filter. (After
70
I
('~
I I
I SUBSTRATE
[2.78])
60 I I
I ,
I ,
I ,
~ 50
II
I ,
,,
~
:>
40
I ,
~ I ,
Id I I
...J
"- I I
1&1 I I
a: 30 I ,
I ,
: I
20 : I
I
I
I
10 I
,'\ I
, \1 I
, \ I
0
4 2 0 • 2 4
WAVELENGTH (A)
z
Fig. 2.31. Side view of a corrugated slab waveguide. Here ho is the average film thickness,
Ll.h the amplitude of the corrugation, and A the period
where heff is the effective guide thickness defined in (2.1.33) and b is the
normalized guide index as explained in Sect. 2.1. We use this result to rewrite
dN / dh as follows
dN dN db (ni- n;)(l - b)
(2.6.47)
dh = db . dh = N heff
82
With this, we can express the coupling coefficient as
7r .dh n; - N2
KTE= - - (2.6.48)
.A heff N
Similar manipulations yield an expression for the coupling coefficient
of the TM modes, i.e.
7r .dh n; - N 2
KTM =- - p (2.6.49)
.A heff N
where p is a reduction factor defined by
(N /nf? -- (N /ne)2 + 1
(2.6.50)
p = (N /nf)2 + (N /ne)2 - 1
Coupled-Mode Formalism
To derive the same coupling coefficients with the more general coupled-mode
formalism, we consider the perturbations caused by the corrugation of the
guide and use the modal fields of the slab guide tabulated in Sect. 2.3.
In terms of the refractive indices nf and ne of film and cover, the cor-
rugation produces a perturbation .de: which can be written as
.de: = e:o(n; - n~) for h(z) > ho
.de: = -Eo(n; - n~) for h(z) < ho (2.6.51)
We insert this into (2.6.21 and 22) to determine the coupling coefficients,
restricting the discussion to "backward" scattering from a forward propa-
gating mode (j-t) to a backward propagating mode of the same mode number
( -j-t).
For TE modes we get K~,_jt = 0, and
11 11
+00 +00
(2.6.52)
where we have assumed that .dh is small enough so that we can replace
Ey( x) by the constant field value Ee assumed by the mode at the film-
cladding interface. The relation of K!,_Il(z) to the coupling coefficient K of
the coupled-wave equations is also indicated. From (2.3.11,12) we can get
the normalized value of E e , and determine Kagain as
.dh n; - N 2
7r
K =- - (2.6.53)
.x heff N
where N is the effective index and heff the effective guide thickness as used
83
10- 1
If =1
K
10-2
before. This can be shown to agree with the formula given by Marcuse [2.2]
who also gives results for coupling between modes of unequal mode number.
One remarkable thing about this formula is that neither of the subscripts c
or s for cladding or substrate appear. This indicates that we get exactly the
same coupling coefficient K when the corrugation is made on the substrate-
film interface instead of the film-cover interface. The reason for this may be
found in (2.3.11) which indicates that a smaller value of (n; - n;) is exactly
balanced by a larger value of the field strength Es at the film-substrate
interface.
Figure 2.32, taken from Shank et al. [2.80]' shows the dependence of K
on film thickness h for the example of a GaAIAs waveguide with nf = 3.59,
ns = 3.414 and nc = 1 (solid curves) and nc = 3.294 (dashed curves). The
normalized quantity K: = >'K/(27rL1h) is used as the ordinate. We note that
the K of each mode assumes a maximum value fairly close to cut-off.
To determine the normalized frequency 8, we compare the exponentials
in (2.6.23 and 33) and find
28 = 2f3/J - K (2.6.54)
The scattering is largest at the center frequency where 8 = 0, which corre-
sponds to a center wavelength >'0 and a propagation constant 130 = 27r / >'0
given by the Bragg condition
K = 2130 >'0 = 2NA (2.6.55)
84
Referring to these quantities, we can rewrite 8 in the form
df3
8 = f3J.L - f30 = L1f3 ~ dwL1w = L1w/Vg (2.6.56)
where Llw is the radian frequency deviation from the center frequency, and
Vg is the group velocity of (2.2.80).
For the TM modes there are difficulties which arise in the application
of the coupled-mode formalism to our particular example of a corrugated
waveguide2 • The root of the problem is thought to be a violation of the
boundary conditions of the perturbed guide. This problem has been dis-
cussed in greater detail in [2.83-85].
In addition to these references, the literature contains discussions of
several alternate methods for the analysis of corrugated guides. Examples
are the treatment of a variety of grating profiles in [2.86], the treatment
of gratings of rectangular tooth shape [2.87], the application of Rouard's
method [2.88]' and the analysis of corrugated multi-layer guides [2.89-92].
We have mentioned before that the conversion from TE-to-TM mod~s can
be regarded as another case of a coupled-wave interaction [2.74, 76]. In this
final subsection, we sketch briefly, how this mode conversion is treated with
the coupled-mode formalism presented in this section. We consider TE--to-
TM mode conversion due to a tensor perturbation L1eij (x) of the form
(2.6.57)
2 The author is indebted to D.G. Hall for illuminating discussions on this issue.
85
With this, we can use (2.6.14) to calculate the induced polarization and ob-
tain, after insertion into (2.6.11), for the changes of the complex amplitudes
J
+00
J
+00
J J
-00 +00
J
+00
J
+00
It is easy to see that (2.6.62 and 63) can be transformed into the standard
form (2.6.28 and 29) of the coupled-wave equations by using (2.6.27) and a
normalized frequency deviation 8 of the form
(2.6.67)
which, again, indicates the degree of deviation from synchronism. The mode
conversion problem is nGw cast in a form that allows us to apply directly
the coupled-wave solutions of Sect. 2.6.3. The overlap integrals of (2.6.64-
66) have to be evaluated for each specific case; examples have been given
by Yariv [2.74]' and Sosnowski and Boyd [2.76J.
86
References
2.1 N.S. Kapany, J.J. Burke: Optical Waveguides (Academic, New York 1974)
2.2 D. Marcuse: Theory of Dielectric Optical Waveguides (Academic, New York 1974)
2.3 D. Marcuse: Light TI-ansmission Optics (Van Nostrand Reinhold, New York 1972)
2.4 M.S. Sodha, A.K. Ghatak: Inhomogeneous Optical Waveguides (Plenum, New York
1977)
2.5 H.G. Unger: Planar Optical Waveguides and Fibers (Clarendon, Oxford 1977)
2.6 M.J. Adams: An Introduction to Optical Waveguides (Wiley, Chichester 1981)
2.7 T. Okoshi: Optical Fibers (Academic, New York 1982)
2.8 A.W. Snyder, J.D. Love: Optical Waveguide Theory (Chapman and Hall, London
1983)
2.9 H.A. Haus: Waves and Fields in Optoelectronics (Prentice-Hall, Englewood Cliffs,
NJ 1984)
2.10 W.W. Anderson: IEEE J. QE-1, 228 (1965)
2.11 A. Reisinger: Appl Opt. 12, 1015 (1973)
2.12 I.P. Kaminow, W.L. Mammel, H.P. Weber: Appl. Opt. 13, 396 (1974)
2.13 D.F. Nelson, J. McKenna: J. Appl. Phys. 38,4057 (1967)
2.14 S. Yamamoto, Y. Koyamada, T. Makimoto: J. Appl. Phys. 43,5090 (1972).
2.15 V. Ramaswamy: Appl. Opt. 13, 1363 (1974)
2.16 V. Ramaswamy: J. Opt. Soc. Am. 64, 1313 (1974)
2.17 P.K. Tien: Appl. Opt. 10, 2395 (1971)
2.18 S.J. Maurer, L.B. Felsen: Proc. IEEE 55,1718 (1967)
2.19 H.K.V. Lotsch: Optik 27, 239 (1968)
2.20 H. Kogelnik, V. Ramaswamy: Appl. Opt. 13, 1857 (1974)
2.21 K. Artmann: Ann. Physik 2, 87 (1948)
H.K.V. Lotsch: Optik 32, 116, 189, 299, 553 (1970/71)
2.22 H. Kogelnik; T.P. Sosnowski, H.P. Weber: IEEE J. QE-9, 795 (1973)
2.23 J.J. Burke: Opt. Sci. Newslett. 5, 31 (Univ. Arizona, 1971)
2.24 J. McKenna: Bell Syst. Tech. J. 46, 1491 (1967)
2.25 R.B. Adler:Proc. IRE 40, 339 (1952)
2.26 W.P. Allis, S.J. Buchsbaum, A. Bers: Waves in Anisotropic Plasmas (Wiley, New
York 1962)
2.27 H. Kogelnik, H.P. Weber: J. Opt. Soc. Am. 64,174 (1974)
2.28 L.D. Landau, E.M. Lifshitz: Electrodynamics of Continuous Media (Pergamon,
Oxford 1960)
2.29 L.D. Landau, E.M. Lifshitz: Quantum Mechanics (Pergamon, Oxford 1958)
2.30 A.D.Berk: IRE Trans. AP-4, 104 (1956)
2.31 R.F. Harrington: Time-Harmonic Electromagnetic Fields (McGraw-Hill, New York
1961)
2.32 K. Kurokawa: IRE Trans. MTT-10, 314 (1962)
2.33 K. Moroshita, N. Kumagai: IEEE Trans. MTT-25, 34 (1977)
2.34 M. Matsuhara: J. Opt. Soc. Am. 63, 1514 (1973)
2.35 H.F. Taylor: IEEE J. QE-12, 748 (1976)
2.36 S.K. Korotky, W.J. Minford, L.L. Buhl, M.D. Divino, R.C. Alferness: IEEE J.
QE-18, 1976 (1982)
2.37 M. Geshiro, M. Ohtaka, M. Matsuhara, N. Kumagai: IEEE J. QE-14, 259 (1978)
2.38 H.A. Haus, W.P. Huang, S. Kawakami, N.A. Whitaker: IEEE J. Lightwave Tech-
nology LT-5, 16 (1987)
2.39 S. Akiba, H.A. Haus: Appl. 0pt. 21, 804 (1982)
2.40 M. Geshiro, M. Ohtaka, M. Matsuhara, N. Kumagai: IEEE J. QE-10, 647 (1974)
2.41 Y. Yamamoto, T. Kamiya, H. Yanai: IEEE J. QE-ll, 729 (1975)
2.42 J.N. Poiky, G.I. Mitchell: J. Opt. Soc. Am. 64,.274 (1974)
2.43 G.E. Smith: IEEE J. QE-4, 288 (1968)
2.44 V.V. Cherny, G.A. Juravlev, A.I. Kirpa, I.L. Rylov, V.P. Ijoy: IEEE J. QE-15,
1401 (1979)
2.45 H.C. Casey Jr., M.B. Panish: Heterostructure Lasers A, (Academic, New York
1978)
87
2.46 H. -Kressel, J .K. Butler: Semiconductor Lasers and Heterojunction LEDS (Aca-
demic, New York 1977), p. 137
2.47 M. Born, E. Wolf: Principles of Optics (Pergamon, New York 1959), p. 50
2.48 F. Abeles: Ann. Physique 5, 596 (1950)
2.49 H. Kogelnik, T. Li: Appl. Opt. 5, 1550 (1966)
2.50 G. PoscW, E. Teller: Z. Physik 83, 143 (1933)
2.51 J.P. Gordon: Bell Syst. Tech. J. 45,321 (1966)
2.52 E.M. Conwell: Appl. Phys. Lett. 23, 328 (1973)
2.53 J.R. Carruthers, I.P. Kaminow, L.W. Stulz: Appl. Opt. 13, 2333 (1974)
2.54 R.D. Standley, V. Ramaswamy: Appl. Phys. Lett. 25, 711 (1974)
2.55 H.A.Haus, R.V. Schmidt: Appl. Opt. 15, 774 (1976)
2.56 B.S. Jeffreys: Quantum Theory, Vol. I, ed. by D.R. Bates (Academic, New York
1961)
2.57 L.B. Felsen, N. Marcuvitz: Radiation and Scattering of Waves (Prentice Hall,
Englewood Cliffs, NJ 1973)
2.58 J.E. Goell: Bell Syst. Tech. J. 48, 2133 (1969)
2.59 W. Schlosser, H.G. Ugger: Advances in Mircowaves (Academic, New York 1966)
2.60 W.O. Schlosser: A.E.U. 18, 403 (1964)
2.61 K Ogusu: IEEE Trans. MTT-25, 874 (1977)
2.62 E.A.J. Marcatili: Bell Syst. Tech. J. 48, 2071 (1969)
2.63 A. Kumar, K Thyagarajan, A.K Ghatak: Opt. Lett. 8, 63 (1983)
2.64 R.M. Knox, P.P. Toulios: Proc. MRI Symp. Submillimeter Waves (Polytechnic
Press, Brooklyn, 1970), p. 497
2.65 V. Ramaswamy: Bell Syst. Tech. J. 53,697 (1974)
2.66 G.B. Hocker, W.K Burns: Appl. Opt. 16, 113 (1977)
2.67 W.H. Zachariasen: Theory of X-Ray Diffraction in Crystals (Wiley, New York
1945)
2.68 S.E. Miller: Bell Syst. Tech. J. 33, 661 (1954)
2.69 J.R. Pierce: J. Appl. Phys. 25, 179 (1954)
2.70 W.H. Louisell: Coupled Mode and Parametric Electronics (Wiley, New York 1960)
2.71 C.F. Quate, C.D.W. Wilkinson, D.K. Winslow: Proc. IEEE 53, 1604 (1965)
2.72 H. Kogelnik: Bell Syst. Tech. J. 48, 2909 (1969)
2.73 A.W. Snyder: J. Opt. Soc. Am. 62, 1267 (1972)
2.74 A. Yariv: IEEE J. QE-9, 919 (1973)
2.75 D. Marcuse, IEEE J. QE-9, 958 (1973)
2.76 T.P. Sosnowski, G.D. Boyd: IEEE J. QE-10, 306 (1974)
2.77 I.P. Kaminow: An Introduction to Electrooptic Devices (Academic, New York
1974)
2.78 D.C. Flanders, H. Kogelnik, R.V. Schmidt, C.V. Shank: Appl. Phys. Lett. 24, 194
(1974)
2.79 H. Kogelnik, C.V. Shank: J. Appl. Phys. 43, 2327 (1972)
2.80 C.V. Shank, R.V. Schmidt, B.I. Miller: Appl. Phys. Lett. 25, 200 (1974)
2.81 S.M. Saad: IEEE Trans. MTT-33, 894 (1985)
2.82 P.G. Verly, R. Tremblay, J.W.Y. Lit: J. Opt. Soc. Am. 70, 964 and 1218 (1980)
2.83 W. Streifer, D.R. Scifres, R.D. Burnham: IEEE J. QE-12, 74 (1976)
2.84 G.I. Stegeman, D. Sarid, J.J. Burke, D.G. Hall: J. Opt. Soc. Am. 71,1497 (1981)
2.85 R.W. Gruhlke, D.G. Hall: Appl. Opt. 23, 127 (1984)
2.86 W. Streifer, D.R. Scifres, R.D. Burnham: IEEE J. QE-ll, 867 (1975)
2.87 A. Hardy: IEEE J. QE-20, 1132 (1984)
2.88 L.A. Weller-Brophy, D.G. Hall: J. Opt. Soc. Am. A-4, 60 (1987)
2.89 Y. Yamamoto, T. Kamiya, H. Yanai: IEEE J. QE-14, 245 (1978)
2.90 S.T. Peng, T. Tamir, H.L. Bertoni: IEEE Trans. MTT-23, 123 (1975)
2.91 R. Petit: Electromagnetic Theory of Gratings, Topics Current Phys., Vol. 22 (Springer,
Berlin, Heidelberg 1980)
2.92 KC. Chang, V. Shah, T. Tamir: J. Opt. Soc. Am. 70, 804 (1980)
88