Sei sulla pagina 1di 12

1176

JOURNAL OF LIGHTWAVE TECHNOLOGY, VOL. 26, NO. 9, MAY 1, 2008

Dielectric Waveguide Theory


Christopher R. Doerr, Fellow, IEEE, and Herwig Kogelnik, Life Fellow, IEEE, Fellow, OSA
Invited Paper

AbstractWe attempt a brief and broad overview and historical perspective of the theory of lightwave propagation in dielectric
waveguides. This is a subject that is now documented by extensive literature. As our review task is limited by constraints in both
space and time it will be illustrative rather than comprehensive.
The illustrations are intended to highlight the principal concepts
and theoretical approaches and refer the reader to the literature
for detailed derivations and explanations. Our story celebrates a
theory with a proud history of nearly 100 years, vibrant and rich
in new ideas, and serving as the base for an increasing number of
powerful applications.
Index TermsDielectric materials, dielectic waveguides, optical
ber cables, optical ber communication, optical ber devices,
optical planar waveguides, optical propagation, waveguide bends,
waveguides, waveguide theory.

I. INTRODUCTION
HE TASK of this paper is to give a broad perspective with
historical emphasis on the theory of lightwave propagation in dielectric waveguides. A dielectric is a nonconductor
of electric current. This theory is based on Maxwells equations. In the spirit of this issue, we should recall that these equations were published 147 years ago [1]. Guides of electromagnetic waves were treated by Lord Rayleigh about 37 years later
[2]. For a nonmetallic circular cylindrical dielectric waveguide,
what we now call a ber, solutions for the modal wave propagation were rst obtained in 1910 by Hondros and Debye [3].
Early workers in the eld were mostly interested in longer wavelengths, such as the propagation of microwaves in dielectric
wires and dielectric rod antennas. Examples are the experiments
conducted in 1936 by the microwave waveguide pioneer Southworth [4] at Bell Labs in Holmdel, the predecessor of the current Crawford Hill Labs. By the early 1940s, the classical textbooks on electromagnetic wave propagation by Stratton [5] and
Schelkunoff [6] contained chapters discussing the detailed solutions for the propagating modes of circular dielectric guides
as well as metallic waveguides in terms of Bessel and Hankel
functions. Among the papers they cite is the 1938 publication
by Brillouin on the same subject [7]. It is interesting that the
modal solutions given by Stratton, as well as those by Hondros
[8] and Carson et al. [9], are of very general validity, allowing
the specication of dielectric constants and of nite conductivities (or laser gain) in both core and cladding. These solutions
contain as special cases both the lossless dielectric guides for

Manuscript received December 4, 2007; revised March 13, 2008.


The authors are with Bell Labs, Alcatel-Lucent, Holmdel, NJ 07733 USA
(e-mail: crdoerr@alcatel-lucent.com).
Digital Object Identier 10.1109/JLT.2008.923632

Fig. 1. Low-order mode patterns of a circular ber (and superpositions) from


[10].

zero conductivity, and the hollow metallic waveguides of microwave technology for a lossless core and a cladding of innite
conductivity.
The hollow metal waveguides received a great deal of early
attention for microwave applications such as RADAR and microwave radio transmission. Dielectric waveguides for optical
applications began to receive attention in the late 1950s. These
optical bers were mostly contemplated for imaging tasks and
ber optic face plates. In 1961 Snitzer and Osterberg observed
the mode patterns of bers in the visible [10] as shown in Fig. 1.
Note that they did not yet use a laser for this observation; they
employed a carbon arc and a monochromator.
The advent of the laser, together with the subsequent demonstration of low-loss bers, laid the groundwork for optical
ber communication, a revolution in the telecom industry. This
created an urgent need and interest in the thorough and detailed
understanding of lightwave propagation in the circular guides
of transmission bers and the planar lm and rectangular
guides of semiconductor lasers, optical integrated circuitry, and
guided-wave optoelectronics. The valuable new understanding
of the theory of dielectric waveguides is now covered in the
several thousand pages of about 20 textbooks and book chapters

0733-8724/$25.00 2008 IEEE

DOERR AND KOGELNIK: DIELECTRIC WAVEGUIDE THEORY

1177

Fig. 4. Geometry and coordinates of a (asymmetric) slab guide with a lm


thickness h.
Fig. 2. Geometry and coordinates of a cylindrical dielectric waveguide. The
dielectric constant " is a function of the cross-sectional coordinates x and y
only.

are the symmetrical three-layer planar slab guide and the circularly symmetric ber.
II. SLAB WAVEGUIDES (1-D)
Planar lm or slab waveguides are dielectric guides with
a 1-D cross-section such as shown in Fig. 4. Neither the index
distribution nor the eld amplitudes depend on the - or -coordinates. Slab guides are the relatively easily analyzed starting
point for the design of semiconductor lasers, integrated optical
circuits or guided-wave optoelectronic circuits (PICs). The
gure shows a step-index guide with lm thickness , and a
.
cover layer index of
The guided modes of slab guides are classied as either transverse electric (TE) or transverse magnetic (TM) modes. For TE
modes we have

Fig. 3. Typical ! diagram of a dielectric waveguide for one polarization.

[11][30]. We can refer the reader to the detailed derivations


and the extensive list of references provided in these works.
Here, we follow the notation of [23]. The general geometry
of a cylindrical dielectric guide is shown in Fig. 2 where the coordinate system is also indicated. Light propagates in the -direction. At the angular frequency the propagation constants
and , respectively,
of free space and the guide are
where c is the velocity of light in free space. The complex amand
describe the elds of electromagplitudes
.
netic modes propagating like
The diagram of Fig. 3 shows the typical propagation
characteristics of a dielectric guide for one of the two polarizations. There is a discrete set of guided modes, two of which are
shown, and a continuous spectrum of radiation modes. -values
cannot occur, where
is the refractive index
larger than
of the guiding lm (or ber core).
is the index of
the substrate (or ber cladding). In general, the guided modes
have a cutoff frequency larger than zero and increasing with
mode number. The spectral region between the cutoff of the
fundamental and the rst-order mode allows only single-mode
propagation. It is the region desired for broadband applications.
There are at least two special cases for which analytical results are available (as discussed below) indicating a zero-frequency cutoff for the fundamental guided mode. These cases

and for TM modes

A. Step Index Guides


To broaden the applicability of theoretical results, one uses
the concepts of an effective index

and of a normalized frequency

As indicated in Fig. 3,
indices

is bounded by the substrate and lm

The fundamental guided TE mode has a normalized cutoff frequency of

1178

JOURNAL OF LIGHTWAVE TECHNOLOGY, VOL. 26, NO. 9, MAY 1, 2008

For a symmetrical guide this cutoff is zero,


. The propagation characteristics of this mode are shown in Fig. 7 as a
special case. The cutoff frequency of the th-order mode is

The elds of the guided modes are sinusoidal in the lm, and
decay exponentially in the substrate and cover regions with increasing distance from the lm.
B. Graded Index Guides
Guide fabrication processes such as diffusion or ion implantation [31] lead to graded index proles where the index
varies gradually in the guide. Some of these guides can be approximated by the parabolic prole
corresponding to the potential well of the harmonic oscillator,
characterizing the effective guide width. The fundawith
mental TE mode of this well is a Gaussian eld with a beam
width 2
where
The higher order modes are HermiteGaussian functions, like
the standard modes of laser beams and resonators [32]. The
2-D generalization of the parabolic slab guide is the SELFOC
Fiber introduced by Uchida et al. possessing favorable multimode dispersion characteristics [33].
Solutions are available for the modes of other analytic
graded index proles, among them the exponential prole and
the 1/cosh prole. Several are listed in [23], e.g., approximate solutions for other proles can be obtained with the WKB
method widely used in quantum mechanics (e.g., see [23]).
C. Multilayer Guides
Multilayer guides consisting of more than three layers are
used for a variety of purposes such as tailoring the waveguide dispersion, obtaining phase matching for guided-wave
nonlinear optics, and providing separate connement in semiconductor heterostructure lasers. Multilayers can also be used
to approximate graded index proles. A compact systematic
theory of multilayer guides can be based on the matrix theory
developed in classical optics for the determination of reection
and transmission properties of multilayer stacks. For further
detail, the reader is referred to [34] and [23].
III. CYLINDRICAL WAVEGUIDES (2-D)
Cylindrical waveguides, illustrated in Fig. 1, conne the light
in two dimensions (
). Main representatives are the circular
optical ber used for transmission of signals over long distances,
and the strip or channel guides used in optical devices. We have
discussed the early theoretical treatment of circular dielectric
guides above, and should add the paper by Snitzer [35] focusing
on optical guides. As the existing rigorous solutions for round
optical bers are rather unwieldy, we discuss the key concepts
using the approximation of weakly guiding bers.

Fig. 5. Dispersion diagram for weakly guiding bers, from Gloge [37]. The
displayed mode numbers are the linearly polarized (LP) mode numbers, the rst
digit being the azimuthal mode number and the second being the radial mode
number.

A. Weakly Guiding Fibers


Approximations for weakly guiding circular bers were discussed by Snyder [36] and Gloge [37], [38]. They assume a
only slightly smaller than the index of the
cladding index
ber core
where
Then, the modal solutions are essentially TE and linearly polarized. For a core radius , the electic eld component is
for
for
where are the Bessel functions and
the modied Hankel
functions, is the azimuthal mode number, and , are the
circular coordinates.
is the maximum eld at the core/
cladding interface. The parameters u and w are related by the
matching conditions, the normalized frequency

(which is the same as for a slab waveguide, except , which is


half the ber diameter, is usually replaced by , which is the full
slab thickness) and the effective index , where

Fundamental mode solutions exist without cutoff, and single. However, low fremode operation is assured as long as
are usually avoided as guidance is very
quencies with
weak and bend losses are high in this region. Fig. 5 shows the
normalized dispersion diagram [37] where the normalized propagation constant b is plotted for several low orders as a function
of V. b is dened by
b
The use of the normalized b-parameter makes the numerical
data very broadly applicable.

DOERR AND KOGELNIK: DIELECTRIC WAVEGUIDE THEORY

Fig. 6. Cross sections of six rectangular channel guides.

1179

Fig. 8. Simulated mode patterns for the rectangular guide with w=h = 1. From
Goell [39].

Fig. 9. Top view and cross section of a rib guide illustrating the effective index
method.

Fig. 7. Dispersion diagram for buried channel guides of height h and width w ,
with w=h ratios of 1, 2, and . From Goell [39].

B. Rectangular Waveguides
Channel waveguides are extensively used in complex integrated optics and optoelectronic circuitry. Fig. 6 shows the cross
sections of six different rectangular guide structures in use. Corresponding graded channel guides can be made by diffusion or
ion implantation.
Analytic solutions for rectangular guides are not available except for the 2-D parabolic prole discussed above. One has to
employ numerical simulations to nd modal solutions, some involving the use of variational methods, or use approximation
methods such as the effective index method (Section III-C). The
rst results for buried rectangular channels were obtained by
Goell in 1969 [39] who used cylindrical space harmonics for
his simulation. Fig. 7 shows his normalized dispersion diagram
for buried channels for a selection of widths.
Fig. 8 shows Goells calculated intensity patterns for the
lower order modes of a guide with unity aspect ratio.
C. Effective Index Method
This method of great intuitive appeal traces back to the paper
of
by Knox and Toulios [40]. It uses the effective index

1-D slab guides to determine propagation characteristics of 2-D


channel guides and gain understanding of structures such as
guided-wave prisms, lenses and gratings. Fig. 9 illustrates the
application of the effective index method to a rib guide. The
guide consists of a lm of thickness and width forming
the rib, and two lateral lms of thickness . If these lms were
of innite extent they would form 1-D slab guides of effective
and
respectively. When the structure is viewed
index
from above, i.e., in the -direction, it looks the same as a 1-D
and a latslab guide with a lm of thickness w and index
. The propagation characteristics of
eral cladding of index
this slab guide are determined and used as an approximation for
the rib guide. Comparisons of these with more exact numerical
simulations have shown reasonably good agreement [41].
IV. WAVEGUIDE BENDS
The understanding of waveguide bends is an important and
complex subject. Knowledge of bend losses is critical for designing densely packed optical circuitry and in the ber to the
home installations that are now seeing deployment on a largescale. The literature on curved guides contains a considerable
variety of calculations and concepts, an early contribution being
the 1969 paper by Marcatili [42] on curved slab and rectangular
guides.
Analysis by a conformal transformation, discussed in 1976 by
Heilblum and Harris [43], provides good intuitive understanding

1180

JOURNAL OF LIGHTWAVE TECHNOLOGY, VOL. 26, NO. 9, MAY 1, 2008

Fig. 10. Transformed index prole of a slab guide with a tight bending radius
at the outer wall of 25 m.

Fig. 11. Outward shift of the modal eld in a bent ber for R =
R = 1 cm, right. From [45].

1, left, and

of waveguide bends. Here, the elds and the dielectric distribution of the curved guide are transformed to those of a straight
in a polar coordinate
guide. If the index is represented by
system, then the transformed index with Cartesian coordinate
is

where
can be chosen arbitrarily. The transformed index proequal
le is shown in Fig. 10 for a curved slab guide with
to the radius of the outer wall. It indicates the exponential increase of the cladding index as the distance from the center of
curvature increases. This results in modes with power leakage,
i.e., radiation loss due to bending. Heilblum [44] has pointed out
the close analogy of the transformed index prole of the curved
guide to that of a prism coupler where leakage is induced intentionally by a high prism index. The above transformation also
changes the thickness of the transformed guide relative to that
of the bent guide, an effect that becomes more pronounced for
tighter bends.
A second effect due to bending is a shift of the mode pattern
away from the center of curvature as if due to a centrifugal
force. This is shown in the simulation of Fig. 11 for a bent
ber [45]. The shift, if not taken into account, can lead to
considerable coupling losses when joining bent and straight
guides or bends with different curvature. For tight bends of
slab guides, this centrifugal shift can be so strong that guiding

is accomplished solely by the outer lm/cladding interface, as


pointed out by Sheem and Whinnery [46] in 1975. In this case
we are dealing with a whispering gallery mode of propagation
as described by Lord Rayleigh [47, p. 617] in 1912. Coupling
losses can be minimized by using offsets of the guide axes to
best match the mode patterns of coupled dissimilar guides.
Details of the corresponding analysis and literature for slab
and rectangular guides are given in the recent publications by
Pennings [48] and Smit et al. [49] in addition to the textbooks
listed [11][30]. Most are scalar theories for curved slabs,
subsequently extended to channel guides by techniques such as
the effective index method.
Theories of the bend losses of circular step-index bers in
the mid 1970s include those of Lewin [50], Snyder et al. [51],
and Marcuse [52]. They were scalar analyses assuming weakly
guiding bers and used a variety of approaches. A recent paper
by Smink et al. [45] provides a summary of the literature on ber
bends and uses a fully vectorial approach valid also for smaller
radii of curvature. They have used this approach for a numerical test of the simplied scalar bending-loss formula, nding
good agreement for large radii of curvature, , but considerable
disagreement for small (below 1 cm) where the simplied formula always overestimated the loss. The simplied formula for
the power loss coefcient of bent weakly guiding single-mode
bers is (e.g., see [52])

where the parameters


etc. refer to the unbent ber
and are explained in Section III-A. The factor is

V. GUIDED-WAVE SIMULATION
There exist analytical expressions for the eigenmodes of slab
waveguides and circular waveguides (even though the propagation constant is found by solving a transcendental equation).
Nearly all other waveguide structures require approximations or
numerical techniques.
There are three main categories of numerical waveguide problems. One is normal mode solving, which is to nd the normal
modes (i.e., eigenmodes) of a guiding structure that is independent of . Another is beam propagation, which is to nd the
change in a eld distribution as it propagates in one direction
through a structure changing with , such as guide transitions
or tapers. Here the results of mode solving are often applied
to characterize short local sections in terms of local normal
modes [53]. The third category involves omnidirectional problems with waves propagating in both -directions, such as structures that contain rings, reectors, or scatterers.
A. Normal Mode Solving
A normal mode solver is used to nd the modes and dispersion diagram of a nonaxially varying waveguide, such that the
time dependence is
(i.e., monochromatic waves) and all
portions of the waveguide cross section have the same propagation constant in the -direction. The search is usually limited to
purely real propagation constants. A mode solver can use either

DOERR AND KOGELNIK: DIELECTRIC WAVEGUIDE THEORY

1181

Another BPM is Fourier transform BPM (FT-BPM) [60].


Fourier optics is based on the notion that any optical eld
can be viewed as a sum of plane waves traveling in
different directions

Fig. 12. FEM calculation of the lowest order guided mode in a microstructured
ber [55], like the ber in Fig. 21, for three different normalized propagation
constants.

the approximate scalar form of Maxwells equations or the full


vectorial form.
A simple method to employ is a scalar method related to the
matrix method for solving for modes in dielectric stacks mentioned earlier. It divides the 2-D structure into rectangular regions, represents the mode as a 2-D Fourier series, matches
boundary conditions, and solves the resulting set of equations
by matrix inversion [54].
The most general method is the nite-element method (FEM),
whereby the structure is divided into triangles and a set of equations relating the regions is often solved by matrix inversion
[55]. It can be scalar or vectorial. An early FEM calculation of
the fundamental guided mode of an early microstructured ber
(as in [56]) is shown in Fig. 12 for different values of , resulting
in the three values of the normalized propagation constant b indicated.
A very different mode solving strategy is to use a beam
propagation method (see the next section) along the imaginary
axis [57]. Nonguided modes rapidly dissipate, leaving only
the normal modes. This technique is more suitable than the
previous ones for nding leaky/lossy/amplifying modes, i.e.,
complex propagation constants.
For periodic structures with high index contrast, such as photonic crystals, a commonly used technique for calculating the
dispersion diagram is the plane-wave expansion method [58].
B. Beam Propagation Methods
The beam propagation method (BPM) starts with an optical
eld in one location and calculates how it evolves in space
and/or time. Thus the eld is not guided in at least one dimendepends
sion. BPM assumes that the eld at location
and not on
and thus is not aponly on the eld
plicable to general scattering problems. Whereas mode solving
is an eigenvalue problem, beam propagation is an initial value
problem. In traditional BPM, the time dependence is assumed
.
to be
One BPM is the local normal mode transfer method [59], also
called the eigenmode expansion method (EME). The structure
, and at each step the normal modes
is divided into steps of
of the local cross-section are calculated. One then computes the
overlap of each local normal mode with those of the previous
step, allowing one to keep track the energy in each mode. This
method aids in understanding and optimizing the efciency of
structures, but has difculty dealing with coupling to unguided
modes.

is the angular spectrum of the eld [61] and represents


the plane wave amplitude propagating at angles
and
in the - and -planes, respectively, from the
-axis. If one starts with a eld
at one plane, then to
nd the eld at a distance one takes the Fourier transform of
to get the angular spectrum, multiplies each plane wave
by a phase term corresponding the distance traveled, and then
takes the inverse Fourier transform

To handle an arbitrary distribution of dielectric between the start


and nish, FT-BPM takes a split-step approach. At each step,
the eld is rst phase shifted by the index distribution and then
is Fourier transformed to the plane-wave domain and is propa. It can be
gated assuming a homogenous medium of length
implemented in a few lines of code.
by
.
Step 1) Multiply
.
Step 2) Take the discrete Fourier transform to get
Step 3) Multiply
by
Step 4) Take the inverse discrete Fourier transform to get a
.
new
and go back to step 1).
Step 5) Increase by
FFT-BPM is accurate for large propagation angles, but it is not
accurate for large index contrasts, and it is scalar. It is also too
slow to be included in most waveguide software packages. A
variation of split-step Fourier BPM that is faster by being optimized for step-index waveguides is called sinc BPM [62].
Most commercial BPM packages use nite-difference (FD)
methods. These can be vectorial [63] and compute quickly. They
have trouble handling large propagation angles, however. Pade
approximants have been employed to improve the angle tolerance [64].
C. Omnidirectional Problems
Many structures not only have large propagation angles but
even have backward propagation, such as is encountered in ring
resonators or structures with high contrast features that are small
compared to a wavelength, such as photonic crystals. Furthermore, some structures have nonlinear elements.
One approach to handling such structures is to use EME in a
bidirectional fashion, treating each segment of the structure as a
scattering matrix with forward and backward traveling components [65].
A more general approach is to use the FEM, previously mentioned for use in normal mode solving. By setting the eld to a
value at one boundary, FEM can calculate the eld everywhere
in the structure. Because FEM is not a BPM, where one can stop

1182

JOURNAL OF LIGHTWAVE TECHNOLOGY, VOL. 26, NO. 9, MAY 1, 2008

Fig. 14. Interaction region between two strip guides of a directional coupler.

gation constants of the two guides are matched, we get coupling


and exchange of power between the two guides. Fig. 14 shows
an electron micrograph of the interaction region of a coupler implemented with glass guides [70].
A. Coupled-Mode Theory

Fig. 13. FD-TD simulation of a pulse traveling in a photonic crystal directional


coupler [69].

the problem before it reaches a boundary, the boundaries must


be handled carefully in FEM. Typically perfectly matched layers
(PMLs) are employed. These layers perform the nontrivial task
of absorbing without reecting.
The most general approach is to no longer assume that the
. Two such methods are the nite-diftime dependence is
ference time-domain (FDTD) [66] and nite-element time-domain (FETD) methods [67]. FDTD is the most common of the
two, and it contains no approximations to Maxwells equations.
These methods, which are more general than BPM, are computationally expensive, since for each time step , rst the electric eld and then the magnetic eld in the entire area of interest
must be computed.
To reduce the computation time, if the time spectral bandwidth is small compared to the optical carrier frequency, one can
use a slowly varying envelope approximation. An example of a
pulse propagating in a directional coupler in a photonic crystal
calculated using this time-domain BPM (TD-BPM) [68], [69] is
shown in Fig. 13.
VI. COUPLERS
A considerable variety of coupling devices and structures is
employed in the construction of modern optical integrated circuits. Among them is the directional coupler, a familiar device
from microwave technology. Its implementation with dielectric
waveguides consists of two strip guides approaching each other,
running close and parallel over an interaction distance, and then
separating again. This is illustrated in Fig. 13. When the propa-

Many phenomena occurring in physics and engineering can


be viewed as coupled-mode or coupled-wave processes.
Among them are the diffraction of X-rays in crystals [71],
the microwave directional coupler [72], other microwave
phenomena [73], and the diffraction in thick holograms [74].
Applying these concepts to dielectric guide structures, the coupled-mode theory rst considers the basic uncoupled guides,
free of perturbations, periodic corrugations, or the presence of
which
another guide. These guides have mode patterns
are orthogonal, and propagate without change in amplitude
with corresponding propagation constants . In the presence
of perturbations or of another guide in the vicinity, there is
coupling from one mode to another or from one guide to another. The strength of the coupling is determined by reciprocity
and orthogonality relations. Now the modal amplitudes are no
longer constant on propagation as power is exchanged between
modes or waveguides.
In the case of the directional coupler, one has mode patterns
and
, one for each guide, with corresponding
propagation constants
and . When the guides get close,
the two modes begin to couple. Their complex amplitudes
and
change on propagation, following the coupled-mode
equations

where the prime indicates differentiation with respect to the


, and is a coupling
propagation distance ,
constant determined by the guide geometry (e.g., see [21], [23],
and [75]). When the guides are phase-matched and mode R is
, while
, the coulaunched with amplitude 1,
pling between the guides leads to amplitudes described by

DOERR AND KOGELNIK: DIELECTRIC WAVEGUIDE THEORY

1183

indicating a sinusoidal interchange of energy with a coupling (or


. As the waves and travel
conversion) length
in the same ( ) directions one talks of codirectional coupling.
Contradirectional coupled wave interactions are also of practical interest [21], [23], [75]. Here, the two coupled waves travel
in opposite directions (typically in a single guide). This coupling can be caused by a periodicity in the guide causing backward Bragg scattering from one mode to the other. The coupled wave equations have the same form as those above, except for some sign changes. The boundary conditions for conand
tradirectional coupling are also changed, e.g., to
, where is the interaction length. For this case, assuming phase-matched guides, we get the amplitudes
Fig. 15. Star coupler.

An example of a periodic guide is a corrugated guide as shown


in Fig. 19. This is used in the construction of a corrugated waveguide lter [76]. Other devices using periodic structures are
Hills ber Bragg gratings [77] and the distributed feedback
laser (DFB) [78]. For the DFB laser one has to enter gain in the
coupled wave equations and modify the boundary conditions.
The coupled mode theory also allows the analysis of deand varying phase
vices with varying coupling strength
. One example is the integrated electrooptic
matching
switch using a directional coupler with alternating
[79].
and chirped gratings are used for the
Apodization of
synthesis of desired grating lter characteristics [80], [81]. For
more detail see [21], [23], [75], and [80].
The above-mentioned periodic structures are periodic in
only (1-D). Examples of 2-D structures are the microstructured
bers mentioned in Section VII, where 3-D photonic crystal
structures are also discussed.
B. Supermodes
In the example of the directional coupler discussed above, the
concept of supermodes (also called normal modes) considers
both single-mode guides of the coupler together as a single
structure in which waves propagate [22]. In this case there
would be two guided modes, the fundamental
with
an even distribution in (in the plane of the circuit), and the
. Their propagation constants
rst-order odd mode
are
and . These modes propagate without change in
amplitude. The link to the coupled-mode concept is established
,
by the superpositions
and by the coupling length equaling the beat length between
. Combining the matrix method of solving multilayer waveguides (Section II-C) with the effective index
concept (Section III-C) provides a convenient way to calculate
the propagation constants of supermodes.
Supermodes are crucial to the understanding of adiabatic couplers, in which the shapes of supermodes at one end of a coupler
are transformed along a coupler without coupling between supermodes. Such mode evolution concepts can be used to make
wavelength and fabrication insensitive couplers [82], polarization splitters [83], and polarization rotators [84].
The supermodes of an innite periodic array of waveguides
have been termed Bloch modes [85], the same name applied

Fig. 16. Arrayed waveguide lens.

to electron waves in crystals. Bloch modes in a periodic waveguide array are analogous to plane waves in free space. For example, the Bloch mode propagating at angle zero consists of
equal powers in all the waveguides, all having the same phase.
Bloch modes are useful in understanding the operation of star
star coupler, shown in Fig. 15, couples
couplers [86]. A
waveguides on one side to waveguides on the other. To a
rst-order approximation, the eld along one curved edge of the
center slab is Fourier transformed along the opposite edge.
Each side of the star coupler consists of a converging array of
waveguides. Dummy waveguides are added to make the array
appear innite. When the waveguides are well separated, the
,
Bloch modes exist entirely within the angular region
i.e., the central Brillouin zone, where is the center-to-center
spacing between the waveguides. For a star coupler with 100%
efciency for all ports in the central Brillouin zone, each of
these Bloch modes in the central Brillouin zone must be adiabatically transformed into a plane wave traveling in the same
direction. Thus the waveguides must be brought together gradually so as to minimize scattering of energy into Bloch modes
outside the central Brillouin zone [85]. Several techniques exist
for helping achieve adiabatic behavior, including segmentation
[87] and vertical tapering [88].
Two star couplers can be connected together with equal length
waveguides to form an imaging system: an arrayed-waveguide
lens. In Fig. 16, the eld along line 1 is imaged along line 2.
If the connecting waveguides instead have a linearly increasing path length, one creates an arrayed-waveguide grating
(AWG), commonly used as a wavelength multiplexer or demultiplexer [89][91], shown in Fig. 17.
C. Self-Imaging
The propagation constants
of the modes of a multimode
slab waveguide approximately obey

1184

JOURNAL OF LIGHTWAVE TECHNOLOGY, VOL. 26, NO. 9, MAY 1, 2008

Fig. 19. Bragg grating in a semiconductor guide, as etched. From Tom Koch.

Fig. 17. Arrayed waveguide grating.

A. Photonic Crystals and Metamaterials

Fig. 18. Simulation of self-imaging in a multimode guide.

The higher the index contrast and the higher the , the more
accurate is this approximation. For reasonably low-order modes

From this, one can see that if we start with all the modes in
, all the modes will be in
phase, then after a distance
phase again. This is the basis of the principle of self-imaging in
multimode slab waveguides and can be used to construct multimode interference couplers [92], [93]. Fig. 18 shows a structure
where a single-mode waveguide abruptly widens to a multimode
waveguide for half the self-imaging distance and then abruptly
narrows to a single-mode waveguide again.
While self-imaging is an approximation in a multimode slab
waveguide, it is exact, in the paraxial approximation, in a multimode parabolic prole waveguide (Section II-B). In such a case

[23] and self-imaging occurs at a distance

VII. MICROSTRUCTURED DIELECTRICS


By making structures with features that are smaller than the
optical wavelength one can conne light very tightly or create
articial dielectrics with special properties. Application of these
microstructures to dielectric waveguides opens many opportunities for improved advanced devices and circuits. It also creates new challenges for the theoretical analysis and simulation
of the new guiding structures. An illustrative example of this is
the simulation of a photonic crystal coupler discussed in Section V-C and shown in Fig. 13.

If the microstructure is periodic, it can be viewed as a photonic crystal [94]. There is a rich theoretical literature about
photonic crystals that draws parallels to electronic crystals, i.e.,
solid-state physics. In photonic crystals, the atoms are analogous to regions of high refractive index. A subset of photonic
crystals includes photonic bandgap materials. The intention is
to create a photonic bandgap (an optical frequency range where
light cannot penetrate the crystal) for all directions and polarizations in either 1, 2, or 3 dimensions.
The concepts fundamental to the theory of photonic crystals
are the Brillouin zones and Bloch waves (or Bloch modes). They
are already discussed in Section VI-B as the supermodes of periodic structures important for the understanding of star couplers.
The Floquet-Bloch theorem, used in solid-state physics, states
that the eigenfunctions, or Bloch modes, of the wave equation
in a periodic potential are equal to the product of a plane wave
and a periodic function that has the same period as the crystal
lattice. For instance, for an innite photonic crystal with a cubic
lattice with an edge length of , the electric eld eigenfunctions
in the crystal must take the form [94]

In 1-D, a photonic crystal is simply a Bragg grating, such as is


shown in Fig. 19. Bragg gratings are widely used in commercial
devices, from DFB lasers [78] to chirped ber Bragg gratings,
as mentioned before.
In 2-D planar structures, the most successful photonic crystal
with a complete bandgap is a triangular lattice of holes in a
high-index material. The ratio of the refractive index between
the material and the holes must be greater than approximately
2.5 in order achieve a true photonic bandgap, and thus only
semiconductor materials have been used in 2-D photonic crystals. To make use of a 2-D photonic crystal, the light must be
guided by conventional waveguiding in the third dimension. Unfortunately, because of the folding of the dispersion diagram in a
periodic structure, dispersion lines inevitably cross the light line
(the lower dashed line in Fig. 3). This means there is always
some phase matching to the continuum, resulting in loss. To

DOERR AND KOGELNIK: DIELECTRIC WAVEGUIDE THEORY

1185

Fig. 20. Three-dimensional photonic crystal, from [98].

mitigate this, the lowest loss 2-D photonic crystals are usually
membranes, with air or oxide above and below the 2-D crystal.
Such membranes are usually made of Si because of its strength.
Probably the most practical application of 2-D photonic crystals
today is to increase the light extraction efciency of light-emitting diodes (LEDs) [95].
The gradient of the lines or surfaces in dispersion diagrams
is equal to the group velocity. Because of the folding of the dispersion diagram, negative group velocities with a positive propagation constant are possible, leading to a negative effective
index. A negative effective index leads to effects such as superlensing. Note that this effective index is dened only for a
narrow range of frequency near the origin in reciprocal space,
just like the effective mass in an electronic crystal.
In 3-D, the most successful photonic crystal with a complete
bandgap is a face-center cubic lattice with two different types of
refractive index shape per unit cell. The rst demonstration was
done at microwave frequencies [96]. An optical frequency example is shown in Fig. 20. Three-dimensional photonic crystal
research is still in its infancy, because such crystals have exceedingly complex dispersion diagrams and because they are difcult to fabricate.
Metamaterials are an articial material made of two or more
different structures or materials. A main difference between a
photonic crystal and a metamaterial is that metamaterials often
include nondielectric materials, such as metals. A typical goal of
a metamaterial is to create a negative effective refractive index
for the purpose of making a superlens [97].
B. Microstructured Fibers
Microstructured bers fall into two main categories. The rst
is guiding the light in a solid core with a cladding of very low
index (e.g., air). The core is held to the ber jacket via a microstructure of material. An early example of such a ber, called
a single-material ber or holey ber, is shown in Fig. 21 [56].
The waveguiding principle is the same as in a conventional ber,
except that the index contrast is very high. This permits a very
small mode size, resulting in a high nonlinearity in a short ber

Fig. 21. Fiber with a pure silica core and an air cladding, [56].

Fig. 22. Fiber with an air core and a photonic crystal cladding, from BlazePhotonics, Ltd [100].

length, and is has been used to generate supercontinuum from a


laser pulse train [99]. It also allows for tight bending of the ber,
which is useful for ber routing in circuit packs and buildings.
The second category is guiding the light in a hollow core with
a photonic crystal cladding, as shown in Fig. 22. Such bers
are typically called photonic crystal bers (PCFs). The waveguiding in PCFs is unconventional in that the light is prevented
from escaping the low-index core by a photonic bandgap in the
cladding. [100] These bers often have the opposite purpose:
minimize nonlinearity. Also, there is a hope that PCFs may one
day have a lower loss than conventional ber, because the loss
of conventional ber is dominated by Rayleigh scattering in the
core, an effect not present in a hollow-core PCF.

1186

JOURNAL OF LIGHTWAVE TECHNOLOGY, VOL. 26, NO. 9, MAY 1, 2008

VIII. CONCLUSION
The theory of lightwave propagation in dielectric waveguides
has a proud history of nearly 100 years. As a symbolic celebration of this milestone we provide 100 references selected
from the vast literature covering the eld. We nd the eld
vibrant and rich in new ideas. New sophisticated simulation
techniques are serving as the foundation for the design of integrated circuitry of increasing complexity. New guide structures such as microstructured bers and photonic crystals are
challenging both the theorist and the experimenter. The story
is, of course, even larger than what we managed to cover in the
available space and time. We have not been able to cover interesting subjects such as guided-wave propagation in optical
ber ampliers, in optical modulators, and in polarization-maintaining bers. Other regrettable omissions include polarization
issues, polarization mode dispersion, optical nonlinear propagation, and the interaction between polarization and nonlinearities.
For these omissions we offer our apologies.
REFERENCES
[1] J. C. Maxwell, On physical lines of force, The London, Edinburgh,
and Dublin Philos. Mag. J. Sci., p. 161, 1861, and J. C. Maxwell, A
Treatise on Electricity and Magnetism, Oxford, U.K.: Clarendon, 1873.
[2] Lord Raleigh, On the passage of electric waves through tubes, or the
vibrations of dielectric cylinders, Phil. Mag., vol. 43, pp. 125132,
1897.
[3] D. Hondros and P. Debye, Elektromagnetische Wellen an dielektrischen Draehten, Ann. Physik, vol. 32, p. 465, 1910.
[4] G. C. Southworth, Hyper-frequency wave guides-general considerations and experimental results, Bell Syst. Tech. J., vol. 15, pp. 284309,
Apr. 1936.
[5] J. A. Stratton, Electromagnetic Theory. NewYork: McGraw-Hill,
1941, p. 524.
[6] S. A. Schelkunoff, Electromagnetic Waves. New York: Van Nostrand,
1943, p. 425.
[7] L. Brillouin, Theoretical study of dielectric cables, Elect. Commun.,
Apr. 1938.
[8] D. Hondros, Ueber elektromagnetische Drahtwellen, Ann. Physik,
vol. 30, pp. 905950, 1909.
[9] J. R. Carson, S. P. Mead, and S. A. Schelkunoff, Hyper-frequency
waveguidesMathematical theory, Bell Syst. Tech. J., vol. 15, pp.
310333, Apr. 1936.
[10] E. Snitzer and H. Osterberg, Observed dielectric waveguide modes in
the visible spectrum, J. Opt. Soc. Amer., vol. 51, pp. 499505, May
1961.
[11] N. S. Kapany, Fiber Optics Principles and Applications. New York:
Academic, 1967.
[12] N. S. Kapany and J. J. Burke, Optical Waveguides. New York: Academic, 1972.
[13] D. Marcuse, Theory of Dielectric Waveguides. New York: Academic,
1974.
[14] H. Kogelnik, Theory of dielectric waveguides, in Integrated Optics. Berlin, Germany: Springer-Verlag, 1975.
[15] M. S. Sodha and A. K. Ghatak, Inhomogeneous Optical Waveguides.
New York: Plenum, 1977.
[16] H. G. Unger, Planar Optical Waveguides and Fibers. Oxford, U.K.:
Clarendon, 1977.
[17] D. Marcuse, D. Gloge, and E. A. J. Marcatili, Guiding properties of
bers, in Optical Fiber Telecom. New York: Academic, 1979, p. 37.
[18] M. J. Adams, An Introduction to Optical Waveguides. New York:
Wiley, 1981.
[19] T. Okoshi, Optical Fibers. New York: Academic, 1982.
[20] A. W. Snyder and J. D. Love, Optical Waveguide Theory. London,
U.K.: Chapman and Hall, 1983.
[21] H. A. Haus, Waves and Fields in Optoelectronics. Upper Saddle
River, NJ: Prentice-Hall, 1984.
[22] W. K. Burns and A. F. Milton, Waveguide transitions and junctions,
in Guided-Wave Optoelectronics, 2nd ed. Berlin, Germany: SpringerVerlag, 1990.

[23] H. Kogelnik, Theory of optical waveguides, in Guided-Wave Optoelectronics, 2nd ed. Berlin, Germany: Springer-Verlag, 1990.
[24] R. E. Collins, Field Theory of Guided Waves, 2nd ed. Piscataway, NJ:
IEEE Press, 1991.
[25] C. Vassallo, Optical Waveguide Concepts. New York: Elsevier Science, 1991.
[26] N. J. Cronin, Microwave and Optical Waveguides. London, U.K.:
Inst. Phys. Publishing, 1995.
[27] S. L. Chuang, Physics of Optoelectronic Devices. New York: Wiley,
1995.
[28] M. R. Amersfoort, J. Bos, X. J. M. Leijtens, and H. J. van Weerden,
Design and simulation tools for integrated optics, in Integrated Optical Circuits and Components. New York: Marcel Dekker, 1999.
[29] C. K. Madsen and J. H. Zhao, Optical Filter Design and Analysis: A
Signal Processing Approach. Hoboken, NJ: Wiley, 1999.
[30] K. Okamoto, Fundamentals of Optical Waveguides, 2nd ed. New
York: Academic, 2006.
[31] R. V. Schmidt and I. P. Kaminow, Metal-diffused optical Wwaveguides in LiNbO , Appl. Phys. Lett., vol. 25, pp. 458460, 1974.
[32] H. Kogelnik and T. Li, Laser beams and resonators, Appl. Opt., vol.
5, pp. 15501567, Oct. 1966.
[33] T. Uchida, M. Furukawa, I. Kitano, K. Koizumi, and H. Matsumara,
Optical characteristics of a light-focusing ber guide and its applications, IEEE J. Quantum Electron., vol. QE-6, pp. 606612, 1970.
[34] J. Chilwell and I. Hodgkinson, Thin-lms eld-transfer matrix theory
of planar multilayer waveguides and reection from prism-loaded
waveguides, J. Opt. Soc. Amer. A, vol. 1, pp. 742753, Jul. 1984.
[35] E. Snitzer, Cylindrical dielectric waveguide modes, J. Opt. Soc.
Amer., vol. 51, pp. 491498, May 1961.
[36] A. W. Snyder, Asymptotic expressions for eigenfunctions and eigenvalues of a dielectric or optical waveguide, IEEE Trans. Microw.
Theory Tech., vol. MTT-17, no. 12, pp. 11301138, Dec. 1969.
[37] D. Gloge, Weakly guiding bers, Appl. Opt., vol. 10, pp. 22522258,
Oct. 1971.
[38] D. Gloge, Propagation effects in optical bers, IEEE Trans. Microw.
Theory Tech., vol. MTT-23, no. 1, pp. 106120, Jan. 1975.
[39] J. E. Goell, A circular-harmonic computer analysis of rectangular dielectric waveguide, Bell Syst. Tech. J., vol. 48, pp. 21332160, Sep.
1969.
[40] R. M. Knox and P. P. Toulios, in Proc. MRI Symp. Submillimeter
Waves, 1970, p. 497.
[41] G. B. Hocker and W. K. Burns, Mode dispersion in diffused channel
waveguides by the effective index method, Appl. Opt., vol. 16, pp.
113118, Jan. 1977.
[42] E. A. J. Marcatili, Bends in optical dielectric guides, Bell Syst. Tech.
J., vol. 48, pp. 21032132, Sep. 1969.
[43] M. Heilblum and J. H. Harris, Analysis of curved optical waveguides
by conformal transformation, IEEE J. Quant. Electr., vol. QE-11, pp.
7583, Feb. 1975.
[44] M. Heilblum, The prism coupler and the dielectric bend: Similarities and anomalous behavior, IEEE J. Quant. Electr., vol. QE-12, pp.
463469, Aug. 1976.
[45] R. W. Smink, B. P. de Hon, and A. G. Tijhuis, Bending loss in optical
bers, J. Opt. Soc. Amer. B, vol. 24, pp. 26102618, Oct. 2007.
[46] S. Sheem and J. R. Whinnery, Guiding by single curved boundaries
in integrated optics, Wave Electron., vol. 1, pp. 105116, 1975.
[47] Lord Rayleigh, The problem of the whispering gallery, in Scientic
Papers. Cambridge, U.K.: Cambridge Univ. Press, 1912, vol. 5.
[48] E. C. M. Pennings, Bends in Optical Ridge Waveguides. Delft, The
Netherlands: Delft Univ. Technol., 1990.
[49] M. K. Smit, E. C. M. Pennings, and H. Blok, A normalized approach
to the design of low-loss optical waveguide bends, J. Lightw. Technol.,
vol. 11, no. 11, pp. 17371742, Nov. 1993.
[50] L. Lewin, Radiation from curved dielectric slabs and bers, IEEE
Trans. Microw. Theory Tech., vol. MTT-22, no. 7, pp. 718727, Jul.
1974.
[51] A. W. Snyder, I. White, and D. J. Mitchell, Radiation from bent optical
waveguides, Electron. Lett., vol. 11, pp. 332333, Jul. 1975.
[52] D. Marcuse, Curvature loss formula for optical bers, J. Opt. Soc.
Amer., vol. 66, pp. 216220, Mar. 1976.
[53] W. K. Burns, Normal mode analysis of waveguide devices, IEEE J.
Lightw. Technol., vol. LT-6, no. 6, pp. 10511068, Jun. 1988.
[54] C. H. Henry and B. H. Verbeek, Solution of the scalar wave equation for arbitrarily shaped dielectric waveguides by two-dimensional
Fourier analysis, J. Lightw. Technol., vol. 7, no. 2, pp. 308313, Feb.
1989.

DOERR AND KOGELNIK: DIELECTRIC WAVEGUIDE THEORY

[55] C. Yeh, K. Ha, S. B. Dong, and W. P. Brown, Single-mode optical


waveguides, Appl. Opt., vol. 18, pp. 14901504, 1979.
[56] P. Kaiser and H. W. Astle, Low-loss single-material bers made from
pure fused silica, Bell Syst. Tech. J., vol. 53, pp. 10211039, 1974.
[57] D. Yevick and B. Hermansson, New approach to low loss optical
waveguide, Electron. Lett., vol. 21, pp. 10291030, 1985.
[58] J. D. Joannopoulos, R. D. Meade, and J. N. Winn, Photonic Crystals.
Princeton, NJ: Princeton Univ. Press, 1995.
[59] C. H. Henry and Y. Shani, Analysis of mode propagation in optical
waveguide devices by Fourier expansion, IEEE J. Quantum Electron.,
vol. QE-27, no. 3, pp. 523530, 1991.
[60] M. D. Feit and J. A. Fleck, Light propagation in graded-index optical
bers, Appl. Opt., vol. 17, pp. 39903998, 1978.
[61] J. W. Goodman, Introduction to Fourier Optics. New York: McGrawHill, 1968.
[62] C. R. Doerr, Beam propagation method tailored for step-index waveguides, IEEE Photon. Technol. Lett., vol. 13, no. 2, pp. 130132, Feb.
2001.
[63] W. Huang, C. Xu, S.-T. Chu, and S. K. Chaudhuri, The nite-diffference vector beam propagation method: Analysis and assessment, J.
Lightw. Technol., vol. 10, no. 3, pp. 295305, Mar. 1992.
[64] G. R. Hadley, Wide-angle beam propagation using Pade approximant
operators, Opt. Lett., vol. 17, pp. 14261428, 1992.
[65] G. Sztefka and H. P. Nolting, Bidirectional eigenmode propagation
for large refractive index steps, IEEE Photon. Technol. Lett., vol. 5,
no. 5, pp. 554557, May 1993.
[66] K. S. Yee, Numerical solution of initial boundary value problems
involving Maxwells equations in isotropic media, IEEE Trans.
Antennas Propagat., vol. AP-14, no. 5, pp. 302307, May 1966.
[67] V. F. Rodriguez-Esquerre, M. Kashiba, and H. E. Hernandez-Figueroa,
Finite-element time-domain analysis of 2-D photonic crystal resonant
cavities, IEEE Photon. Technol. Lett., vol. 16, no. 3, pp. 816818, Mar.
2004.
[68] P.-L. Liu, Q. Zhao, and F.-S. Choa, Slow-wave nite-difference beam
propagation method, IEEE Photon. Technol. Lett., vol. 7, pp. 890892,
Aug. 1995.
[69] M. Koshiba, Y. Tsuji, and M. Hikari, Time-domain beam propagation method and its application to photonic crystal circuits, J. Lightw.
Technol., vol. 18, no. 1, pp. 102110, Jan. 2000.
[70] J. E. Goell, Electron-resist fabrication of bends and couplers for integrated optical circuits, Appl. Opt., vol. 12, pp. 729736, Apr. 1973.
[71] W. H. Zachariasen, Theory of X-Ray Diffraction in Crystals. New
York: Wiley, 1945.
[72] S. E. Miller, Coupled wave theory and waveguide applications, Bell
Syst. Tech. J., vol. 33, pp. 661720, May 1954.
[73] J. R. Pierce, Coupling of modes of propagation, J. Appl. Phys., vol.
25, pp. 179183, Feb. 1954.
[74] H. Kogelnik, Coupled wave theory for thick hologram gratings, Bell
Syst. Tech. J., vol. 48, pp. 29092947, Nov. 1969.
[75] A. Yariv, Optical Electronics, 4th ed. Philadelphia, PA: Saunders
College Publishing, 1991.
[76] D. C. Flanders, H. Kogelnik, R. V. Schmidt, and C. V. Shank, Grating
lters for thin-lm optical waveguide, Appl. Phys. Lett., vol. 24, pp.
194196, Feb. 1974.
[77] K. O. Hill, Y. Fujii, D. C. Johnson, and B. S. Kawasaki, Photosensitivity in optical waveguides: Application to reection lter fabrication,
Appl. Phys. Lett., vol. 32, p. 647, May 1978.
[78] H. Kogelnik and C. V. Shank, Coupled wave theory of distributed
feedback lasers, J. Appl. Phys., vol. 43, pp. 23272335, May 1972.
[79] R. V. Schmidt and H. Kogelnik, Electrooptically switched coupler
with stepped reversal using Ti diffused LiNbO waveguides,
Appl. Phys. Lett., vol. 28, pp. 503506, May 1976.
[80] M. Matsuhara, K. O. Hill, and A. Watanabe, Optical-waveguide lters:
Synthesis, J. Opt. Soc. Amer., vol. 65, pp. 804809, Jul. 1975.
[81] P. S. Cross and H. Kogelnik, Sidelobe suppression in corrugatedwaveguide lters, Opt. Lett., vol. 1, pp. 4345, Jul. 1977.
[82] R. Adar, C. H. Henry, R. F. Kazarinov, R. C. Kistler, and G. R. Weber,
Adiabatic 3-dB couplers, lters, and multiplexers made with silica
waveguides on silicon, J. Lightw. Technol., vol. 10, no. 1, pp. 4650,
Jan. 1992.
[83] Y. Shani, C. H. Henry, R. C. Kistler, R. F. Kazarinov, and K. J.
Orlowsky, Integrated optic adiabatic polarization splitter on silicon,
Appl. Phys. Lett., vol. 56, no. 2, pp. 120121, 1990.
[84] M. R. Watts and H. A. Haus, Integrated mode-evolution-based polarization rotators, Opt. Lett., vol. 30, pp. 138140, Jan. 2005.
[85] C. Dragone, Optimum design of a planar array of tapered waveguides, J. Opt. Soc. Amer. A, vol. 7, pp. 20812093, Nov. 1990.

1187

[86] C. Dragone, Efcient N N star coupler based on Fourier optics,


Electron. Lett., vol. 24, pp. 942944, 1988.
[87] Y. P. Li, Optical device having low insertion loss, U.S. Patent
5 745 618, Apr. 28, 1998.
[88] A. Sugita, A. Kaneko, K. Okamoto, M. Itoh, A. Himeno, and Y.
Ohmori, Very low insertion loss arrayed-waveguide grating with
vertically tapered waveguides, IEEE Photon. Technol. Lett., vol. 12,
no. 9, pp. 11801182, Sep. 2000.
[89] M. K. Smit, New focusing and dispersive planar component based on
an optical phased array, Electron. Lett., vol. 24, pp. 385386, 1988.
[90] H. Takahashi, S. Suzuki, K. Kato, and I. Nishi, Arrayed-waveguide
grating for wavelength division multi/demultiplexer with nanometer
resolution, Electron. Lett., vol. 26, pp. 8788, 1990.
[91] C. Dragone, An N N optical multiplexer using a planar arrangement
of two star couplers, IEEE Photon. Technol. Lett., vol. 3, no. 9, pp.
812815, Sep. 1991.
[92] R. Ulrich and T. Kamiya, Resolution of self-images in planar optical
waveguides, J. Opt. Soc. Amer., vol. 68, pp. 583592, 1978.
[93] L. B. Soldano and E. C. M. Pennings, Optical multi-mode interference
devices based on self-imaging: Principles and applications, J. Lightw.
Technol., vol. 13, no. 4, pp. 615627, Apr. 1995.
[94] J.-M. Lourtioz, H. Benisty, V. Berger, J.-M. Gerard, D. Maystre, and
A. Tchelnokov, Photonic Crystals: Towards Nanoscale Photonic Devices. Berlin, Germany: Springer, 2005.
[95] M. Rattier, H. Benisty, R. P. Stanley, J.-F. Carlin, R. Houdre, U.
Oesterle, C. J. M. Smith, C. Weisbuch, and T. F. Krauss, Toward ultrahigh-efciency aluminum oxide microcavity light-emitting diodes:
Guided mode extraction by photonic crystals, IEEE J. Sel. Top.
Quant. Electron, vol. 8, no. 1, pp. 238247, Mar./Apr. 2002.
[96] E. Yablonovitch, T. J. Gmitter, and K. M. Leung, Photonic band structures: The face-centered-cubic case employing non-spherical atoms,
Phys. Rev. Lett., vol. 67, pp. 22952298, 1991.
[97] J. B. Pendry, Negative refraction makes a perfect lens, Phys. Rev.
Lett., vol. 85, pp. 39663969, 2000.
[98] K. Aoki, H. T. Miyazaki, H. Hirayama, K. Inoshita, T. Baba, K.
Sakoda, N. Shinya, and Y. Aoyagi, Microassembly of semiconductor
three-dimensional photonic crystals, Nature Mater., vol. 2, pp.
117121, Feb. 2003.
[99] J. K. Ranka, R. S. Windeler, and A. J. Stentz, Visible continuum generation in air-silica microstructure optical bers with anamolous dispersion at 800 nm, Opt. Lett., vol. 25, pp. 2527, 2000.
[100] J. C. Knight, J. Broeng, T. A. Birks, and P. St. J. Russell, Photonic
bandgap guidance in optical bers, Science, vol. 282, pp. 14761478,
1998.

Christopher R. Doerr (M97SM06F07) received the B.S. degree in aeronautical engineering


and the B.S., M.S., and Ph.D. degrees in electrical
engineering from the Massachusetts Institute of
Technology (MIT), Cambridge.
He attended MIT on an Air Force scholarship and
earned pilot wings in 1991. Since coming to Bell
Labs, Holmdel, NJ, in 1995, his research has focused
on integrated devices for optical communication. He
was promoted to Distinguished Member of Technical
Staff in 2000.
Dr. Doerr received the OSA Engineering Excellence Award in 2002. He is
the Editor of IEEE PHOTONICS TECHNOLOGY LETTERS.

Herwig Kogelnik (M61F73LF98) was born in


Graz, Austria, in 1932. He received the Dipl.Ing. and
Dr.Tech. degrees in electrical engineering from the
Technical University of Vienna, Austria, in 1955 and
1958, respectively, and the D.Phil. degree in physics
from Oxford University, Oxford, U.K. in 1960.
In 1961, he joined the research division of Bell
Laboratories (now Alcatel-Lucent), Holmdel, NJ,
and is currently Adjunct Photonics Research Vice
President.
Dr. Kogelnik is currently a Fellow and past President (1989) of the Optical Society of America (OSA), as well as a Member of
the National Academy of Science and the National Academy of Engineering.
He is an Honorary Fellow of St. Peters College, Oxford, U.K.

Potrebbero piacerti anche